02.12.2012 Views

References

References

References

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.

Soil Biology<br />

Series Editor: Ajit Varma 9


Volumes published in the series<br />

Volume 1<br />

A. Singh, O.P. Ward (Eds.)<br />

Applied Bioremediation and Phytoremediation<br />

2004<br />

Volume 2<br />

A. Singh, O.P. Ward (Eds.)<br />

Biodegradation and Bioremediation<br />

2004<br />

Volume 3<br />

F. Buscot, A. Varma (Eds.)<br />

Microorganisms in Soils: Roles in Genesis and Functions<br />

2005<br />

Volume 4<br />

S. Declerck, D.-G. Strullu, J.A. Fortin (Eds.)<br />

In Vitro Culture of Mycorrhizas<br />

2005<br />

Volume 5<br />

R. Margesin, F. Schinner (Eds.)<br />

Manual for Soil Analysis –<br />

Monitoring and Assessing Soil Bioremediation<br />

2005<br />

Volume 6<br />

H. König, A. Varma (Eds.)<br />

Intestinal Microorganisms of Termites<br />

and Other Invertebrates<br />

2006<br />

Volume 7<br />

K.G. Mukerji, C. Manoharachary, J. Singh (Eds.)<br />

Microbial Activity in the Rhizosphere<br />

2006<br />

Volume 8<br />

P. Nannipieri, K. Smalla (Eds.)<br />

Nucleic Acids and Proteins in Soil<br />

2006


Barbara J.E. Schulz<br />

Christine J.C. Boyle<br />

Thomas N. Sieber (Eds.)<br />

Microbial Root<br />

Endophytes<br />

With 29 Figures, 4 in Color<br />

123


PD Dr. Barbara J. E. Schulz<br />

Technical University of Braunschweig<br />

Institute of Microbiology<br />

Spielmannstraße 7<br />

38106 Braunschweig<br />

Germany<br />

e-mail: b.schulz@tu-bs.de<br />

Dr. Thomas N. Sieber<br />

Swiss Federal Institute of Technology<br />

Department of Environmental Sciences<br />

Institute of Integrative Biology<br />

Forest Pathology and Dendrology<br />

8092 Zürich<br />

Switzerland<br />

e-mail: thomas.sieber@env.ethz.ch<br />

Library of Congress Control Number: 2005938057<br />

Dr. Christine J. C. Boyle<br />

Augustastraße 32<br />

02826 Görlitz<br />

Germany<br />

e-mail: c.boyle@tu-bs.de<br />

ISSN 1613-3382<br />

ISBN-10 3-540-33525-0 Springer Berlin Heidelberg New York<br />

ISBN-13 978-3-540-33525-2 Springer Berlin Heidelberg New York<br />

This work is subject to copyright. All rights reserved, whether the whole or part of the<br />

material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,<br />

recitation, broadcasting, reproduction on microfilm or in any other way, and storage in data<br />

banks. Duplication of this publication or parts thereof is permitted only under the provisions<br />

of the German Copyright Law of September 9, 1965, in its current version, and permission<br />

for use must always be obtained from Springer. Violations are liable for prosecution under<br />

the German Copyright Law.<br />

Springer is a part of Springer Science + Business Media<br />

springer.com<br />

© Springer-Verlag Berlin Heidelberg 2006<br />

Printed in Germany<br />

The use of general descriptive names, registered names, trademarks, etc. in this publication<br />

does not imply, even in the absence of a specific statement, that such names are exempt<br />

from the relevant protective laws and regulations and therefore free for general use.<br />

Editor: Dr. Dieter Czeschlik, Heidelberg, Germany<br />

Desk Editor: Dr. Jutta Lindenborn, Heidelberg, Germany<br />

Cover design: design&production, Heidelberg, Germany<br />

Typesetting and production: LE-TEX Jelonek, Schmidt & Vöckler GbR, Leipzig, Germany<br />

31/3150-YL - 5 4 3 2 1 0 - Printed on acid-free paper


Preface<br />

Healthy plant roots are not only colonized by mycorrhizal fungi and rhizobial<br />

bacteria, but also by a myriad of other microorganisms, including<br />

endophytic bacteria and fungi. Comparatively little is known about these<br />

endophytic microorganisms, which do not cause apparent disease, but<br />

colonize root tissues inter- and/or intracellulary. Although there had been<br />

previous research studying both bacterial and fungal endophytes, it was<br />

in the mid-1980s that numerous investigators began studying these groups<br />

of microorganisms more intensively. Initially, most work on endophytes<br />

centered on the diversity of isolates and correlations with ecological factors.Recentlyithasbecomeclearthatsomeoftheseinteractionswith<br />

endophytic bacteria and fungi can be latently pathogenic and/or mutualistic.<br />

In mutualistic interactions, the endophyte may improve growth of the<br />

host, convey stress tolerance, induce systemic resistance, or supply the host<br />

with nutrients. On the other hand, most endophytes are also able to grow<br />

saprotrophically, e.g., from surface-sterilized tissues on media containing<br />

dead organic substrates. Thus, it has become obvious that endophytes have<br />

multiple life history strategies and that these can be extremely plastic, as<br />

will become clear to the readers of the subsequent 19 chapters.<br />

This book is the first to deal with bacterial and fungal root endophytes,<br />

their diversity, life history strategies, interactions, applications in agriculture<br />

and forestry, and also with methods for isolation, cultivation, and both<br />

conventional and molecular methods for identification and detection. The<br />

first chapter deals with the question: What are endophytes? However, it<br />

also introduces the reader to the subjects treated in the subsequent chapters.<br />

We hope that readers will not only find this book informative, but<br />

will also be provoked to further study these fascinating interactions, and<br />

in particular to better understand the mechanisms regulating them. It will<br />

become apparent that we are still far from understanding the factors that<br />

determine whether a plant-microbial interaction remains asymptomatic,<br />

leads to disease, or is mutualistic.


VI Preface<br />

We would like to thank our colleagues for their contributions and their<br />

work to make this book a successful unity, to Jutta Lindenborn of Springer<br />

for her friendly help and advice, and to Ajit Varma for the invitation to edit<br />

abookinthisseries.<br />

Braunschweig, Barbara Schulz<br />

Görlitz and Zürich, Christine Boyle<br />

June 2006 and Thomas Sieber


Contents<br />

1 WhatareEndophytes? 1<br />

Barbara Schulz, Christine Boyle<br />

1.1 Introduction and Definitions ............................................ 1<br />

1.2 Colonisation .................................................................. 2<br />

1.3 Assemblages and Adaptation ............................................ 4<br />

1.4 Life History Strategies ..................................................... 6<br />

1.5 Balanced Antagonism...................................................... 7<br />

1.6 Conclusions ................................................................... 9<br />

<strong>References</strong> ............................................................................. 10<br />

Part I Endophytic Bacteria<br />

2 Spectrum and Population Dynamics<br />

of Bacterial Root Endophytes 15<br />

Johannes Hallmann, Gabriele Berg<br />

2.1 Introduction .................................................................. 15<br />

2.2 Population Density ......................................................... 15<br />

2.3 Bacterial Spectrum ......................................................... 16<br />

2.4 Bacterial Diversity .......................................................... 21<br />

2.5 Factors Influencing Colonisation....................................... 21<br />

2.5.1 Methodology....................................................... 21<br />

2.5.2 Geography .......................................................... 22<br />

2.5.3 Plant Species ....................................................... 22<br />

2.5.4 Plant Genotype .................................................... 23<br />

2.6 Interactions ................................................................... 24<br />

2.6.1 Plant Pathogens ................................................... 24<br />

2.6.2 Plant Symbionts................................................... 25<br />

2.6.3 Plant Defence Mechanisms .................................... 25<br />

2.6.4 Agricultural Practices ........................................... 25<br />

2.7 Potential Human Pathogens Among Root Endophytes.......... 26<br />

2.8 Conclusions ................................................................... 27<br />

<strong>References</strong> ............................................................................. 28


VIII Contents<br />

3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 33<br />

Joseph W. Kloepper, Choong-Min Ryu<br />

3.1 Introduction and Terminology ......................................... 33<br />

3.2 Scope of Endophytes that Elicit Induced Resistance<br />

and Pathosystems Affected............................................... 34<br />

3.3 Internal Colonization of Endophytes<br />

that Elicit Induced Resistance ........................................... 39<br />

3.4 Plant Responses to Endophytic Elicitors............................. 41<br />

3.5 Implementation in Production Agriculture:<br />

Two Case Studies ............................................................ 44<br />

3.6 Conclusions ................................................................... 49<br />

<strong>References</strong> ............................................................................. 50<br />

4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 53<br />

Gabriele Berg, Johannes Hallmann<br />

4.1 Introduction .................................................................. 53<br />

4.2 Spectrum of Indigenous Endophytic Bacteria<br />

with Antagonistic Potential Towards Fungal Plant Pathogens 54<br />

4.3 Mode of Action of Antagonistic Bacteria ............................ 58<br />

4.3.1 Antibiosis ........................................................... 58<br />

4.3.2 Competition........................................................ 59<br />

4.3.3 Lysis................................................................... 59<br />

4.3.4 Induction of Plant Defence Mechanisms .................. 60<br />

4.3.5 Plant Growth ....................................................... 60<br />

4.4 Control Potential of Endophytic Bacteria............................ 60<br />

4.5 Enhancing Biocontrol Efficiency ....................................... 61<br />

4.6 Conclusions ................................................................... 65<br />

<strong>References</strong> ............................................................................. 66<br />

5 Role of Proteins Secreted by Rhizobia in Symbiotic<br />

Interactions with Leguminous Roots 71<br />

Maged M. Saad, William J. Broughton, William J. Deakin<br />

5.1 Introduction .................................................................. 71<br />

5.2 Bacterial Protein Secretion Systems................................... 73<br />

5.2.1 Type I Secretion Systems ....................................... 73<br />

5.2.2 Type II Secretion Systems ...................................... 77<br />

5.2.3 Type III Secretion Systems..................................... 77<br />

5.2.4 Type IV Secretion Systems..................................... 82<br />

5.3 Conclusions ................................................................... 83<br />

<strong>References</strong> ............................................................................. 83


Contents IX<br />

6 Research on Endophytic Bacteria: Recent Advances<br />

with Forest Trees 89<br />

Richa Anand, Leslie Paul, Chris Chanway<br />

6.1 Introduction .................................................................. 89<br />

6.2 Bacterial Endophytes of Forest Trees.................................. 91<br />

6.3 Endophytic Bacteria of Conifers........................................ 92<br />

6.4 Modes and Sites of Entry.................................................. 95<br />

6.5 Mechanisms of Plant Growth Promotion............................ 97<br />

6.6 Future Work................................................................... 102<br />

<strong>References</strong> ............................................................................. 103<br />

Part II Endophytic Fungi<br />

7 Biodiversity of Fungal Root-Endophyte Communities<br />

and Populations, in Particular of the Dark Septate Endophyte<br />

Phialocephala fortinii s. l. 107<br />

Thomas N. Sieber, Christoph R. Grünig<br />

7.1 Introduction .................................................................. 107<br />

7.2 Species Diversity of Root Endophyte Communities.............. 108<br />

7.2.1 Geography and Climate......................................... 109<br />

7.2.2 Soil .................................................................... 114<br />

7.2.3 Multitrophic Interactions ...................................... 115<br />

7.2.4 Natural and Anthropogenic Disturbances ................ 117<br />

7.3 Dark Septate Endophytes................................................. 119<br />

7.3.1 History ............................................................... 119<br />

7.3.2 Biodiversity......................................................... 119<br />

7.3.3 Diversity of Phialocephala fortinii........................... 121<br />

7.4 Conclusions ................................................................... 125<br />

<strong>References</strong> ............................................................................. 126<br />

8 Endophytic Root Colonization by Fusarium Species:<br />

Histology, Plant Interactions, and Toxicity 133<br />

CharlesW.Bacon,IdaE.Yates<br />

8.1 Introduction .................................................................. 133<br />

8.2 Plant and Fungus Interactions .......................................... 134<br />

8.2.1 Hemibiotrophic Characteristics.............................. 138<br />

8.2.2 Histology ............................................................ 139<br />

8.2.3 Mycotoxins.......................................................... 143<br />

8.2.4 Mycotoxins and Host Relationships......................... 144<br />

8.2.5 Physiological Interactions and Defense Metabolites... 145<br />

8.3 Summary....................................................................... 146<br />

<strong>References</strong> ............................................................................. 147


X Contents<br />

9 Microbial Endophytes of Orchid Roots 153<br />

Paul Bayman, J. Tupac Otero<br />

9.1 Introduction .................................................................. 153<br />

9.2 Habits and Types of Orchid Roots ..................................... 153<br />

9.3 Bacteria as Epiphytes and Endophytes of Orchid Roots ........ 154<br />

9.4 Orchid Endophytes or Orchid Mycorrhizal Fungi? ............... 155<br />

9.5 Problems with the Taxonomy of Orchid Endophytic Fungi.... 157<br />

9.6 Host Specificity of Orchid Endophytes ............................... 158<br />

9.7 Endophytic Fungi in Roots of Terrestrial,<br />

Photosynthetic Orchids ................................................... 158<br />

9.8 Endophytic Fungi in Roots of Myco-Heterotrophic Orchids .. 167<br />

9.9 Endophytic Fungi in Roots of Epiphytic<br />

and Lithophytic Orchids .................................................. 169<br />

9.10 Endophytic Fungi in Epiphytic Orchid Roots:<br />

Importance to Plant Hosts................................................ 171<br />

9.11 Conclusions ................................................................... 172<br />

<strong>References</strong> ............................................................................. 173<br />

10 Fungal Endophytes in Submerged Roots 179<br />

Felix Bärlocher<br />

10.1 Introduction .................................................................. 179<br />

10.2 Aquatic Hyphomycetes .................................................... 180<br />

10.3 Fungi in Submerged Roots ............................................... 181<br />

10.4 Conclusions and Outlook................................................. 186<br />

<strong>References</strong> ............................................................................. 188<br />

11 Nematophagous Fungi as Root Endophytes 191<br />

Luis V. Lopez-Llorca, Hans-Börje Jansson, José Gaspar Maciá Vicente,<br />

Jesús Salinas<br />

11.1 Introduction .................................................................. 191<br />

11.2 Nematophagous Fungi..................................................... 191<br />

11.2.1 Nematode Parasites .............................................. 192<br />

11.2.2 Mycoparasites...................................................... 195<br />

11.2.3 Root Endophytes.................................................. 195<br />

11.3 Concluding Remarks....................................................... 202<br />

<strong>References</strong> ............................................................................. 203<br />

12 Molecular Diversity and Ecological Roles of Mycorrhiza-Associated<br />

Sterile Fungal Endophytes in Mediterranean Ecosystems 207<br />

Mariangela Girlanda, Silvia Perotto, Anna Maria Luppi<br />

12.1 Introduction .................................................................. 207


Contents XI<br />

12.2 Diversity of DSE Associates of Ecto- and Endo-Mycorrhizal<br />

Plants in Mediterranean Ecosystems in Northern Italy ......... 209<br />

12.3 Ecological Relationships with Conventional Mycorrhizal<br />

and Pathogenic Symbionts ............................................... 214<br />

12.4 Conclusions ................................................................... 219<br />

<strong>References</strong> ............................................................................. 220<br />

13 Oidiodendron maius: Saprobe in Sphagnum Peat,<br />

Mutualist in Ericaceous Roots? 227<br />

Adrianne V. Rice, Randolph S. Currah<br />

13.1 Introduction .................................................................. 227<br />

13.2 Oidiodendron maius as a Saprobe...................................... 230<br />

13.3 Ericoid Mycorrhizas........................................................ 234<br />

13.4 Oidiodendron maius as an Ericoid Mycorrhizal Fungus ........ 237<br />

13.5 Significance and Relevance............................................... 239<br />

13.6 Conclusions ................................................................... 241<br />

<strong>References</strong> ............................................................................. 242<br />

14 Mycorrhizal and Endophytic Fungi of Epacrids (Ericaceae) 247<br />

John W.G. Cairney<br />

14.1 Introduction .................................................................. 247<br />

14.2 Endophytes of Epacrid Roots............................................ 248<br />

14.3 Diversity and Spatial Distribution<br />

of Endophyte Taxa in Epacrid Root Systems........................ 251<br />

14.4 Mycorrhizal Status of Endophytes from Epacrids................. 253<br />

14.5 Saprotrophic Potential of Mycorrhizal Fungi....................... 254<br />

14.6 Symbiotic Functioning of Mycorrhizal Fungi ...................... 255<br />

14.7 Conclusions ................................................................... 257<br />

<strong>References</strong> ............................................................................. 257<br />

15 Mutualistic Interactions with Fungal Root Endophytes 261<br />

Barbara Schulz<br />

15.1 Introduction .................................................................. 261<br />

15.2 Colonisation and Histology.............................................. 262<br />

15.3 Secondary Metabolites..................................................... 264<br />

15.4 Growth Enhancement...................................................... 265<br />

15.5 Disease Suppression........................................................ 269<br />

15.6 Stress Tolerance.............................................................. 271<br />

15.7 Factors Determining the Status of the Interaction................ 272<br />

15.8 Conclusions ................................................................... 273<br />

<strong>References</strong> ............................................................................. 276


XII Contents<br />

16 Understanding the Roles of Multifunctional Mycorrhizal<br />

and Endophytic Fungi 281<br />

Mark C. Brundrett<br />

16.1 How Mycorrhizal Fungi Differ from Endophytes ................. 281<br />

16.1.1 Definitions .......................................................... 281<br />

16.1.2 Roles of Endophytes and Mycorrhizal Fungi............. 282<br />

16.2 Endophytic Activity by Mycorrhizal Fungi.......................... 283<br />

16.2.1 Glomeromycotan (Vesicular-Arbuscular<br />

Mycorrhizal) Fungi............................................... 283<br />

16.2.2 Ectomycorrhizal Fungi.......................................... 286<br />

16.2.3 Fungi in Orchids .................................................. 287<br />

16.2.4 Ericoid Mycorrhizal Fungi ..................................... 290<br />

16.2.5 Endophytic Fungi in Mycorrhizal Roots................... 290<br />

16.3 Issues with the Identification and Categorisation<br />

of Fungi in Roots ............................................................ 291<br />

16.4 Evolution of Root-Fungus Associations.............................. 292<br />

16.5 Conclusions ................................................................... 293<br />

<strong>References</strong> ............................................................................. 293<br />

Part III Methods<br />

17 Isolation Procedures for Endophytic Microorganisms 299<br />

Johannes Hallmann, Gabriele Berg, Barbara Schulz<br />

17.1 Introduction .................................................................. 299<br />

17.2 Surface Sterilisation ........................................................ 300<br />

17.2.1 Pre-treatment ...................................................... 300<br />

17.2.2 Sterilising Agents ................................................. 301<br />

17.2.3 Surfactants.......................................................... 305<br />

17.2.4 Rinsing............................................................... 305<br />

17.2.5 Sterility Check and Optimisation............................ 305<br />

17.3 Culture of Tissue and Plant Fluid of Sterilised Roots<br />

on Nutrient Medium ....................................................... 306<br />

17.3.1 Segments ............................................................ 306<br />

17.3.2 Maceration of Root Tissue ..................................... 306<br />

17.3.3 Centrifugation of Root Tissue ................................ 307<br />

17.4 Vacuum and Pressure Extraction ...................................... 308<br />

17.5 Media............................................................................ 309<br />

17.5.1 Media for Isolating Bacteria................................... 310<br />

17.5.2 Media for Isolating Fungi ...................................... 310<br />

17.5.3 Supplements........................................................ 310<br />

17.5.4 Selective Media .................................................... 311


Contents XIII<br />

17.6 Cultivation-Independent Methods..................................... 311<br />

17.7 Quantification of Colonisation.......................................... 313<br />

17.8 Conclusions ................................................................... 313<br />

<strong>References</strong> ............................................................................. 314<br />

18 Microbial Interactions with Plants: a Hidden World? 321<br />

Guido V. Bloemberg, Margarita M. Camacho Carvajal<br />

18.1 Introduction .................................................................. 321<br />

18.2 Microscopic Techniques for Studying<br />

Plant-Microbe Interactions .............................................. 322<br />

18.2.1 Light Microscopy and Enzymatic Reporters ............. 322<br />

18.2.2 Scanning Electron Microscopy ............................... 324<br />

18.2.3 Epifluorescence Microscopy<br />

and the Application of Auto-Fluorescent Proteins...... 325<br />

18.3 Visualisation of Bacterium-Plant Interactions ..................... 327<br />

18.4 Most Recent Developments in Visualising<br />

Plant-Microorganism Interactions..................................... 330<br />

18.5 Visualisation of Plant-Fungus Interactions ......................... 331<br />

18.6 Future Perspectives ......................................................... 333<br />

<strong>References</strong> ............................................................................. 333<br />

19 Application of Molecular Fingerprinting Techniques<br />

to Explore the Diversity of Bacterial Endophytic Communities 337<br />

LeoS.vanOverbeek,JimvanVuurde,JanD.vanElsas<br />

19.1 Introduction .................................................................. 337<br />

19.2 Colonisation by Bacterial Endophytes................................ 338<br />

19.3 Shifts of Bacterial Endophyte Communities........................ 338<br />

19.4 Molecular methods to Study Bacterial Endophytes .............. 340<br />

19.5 Molecular Fingerprinting of Endophyte Communities.......... 341<br />

19.5.1 Basic Concept of Molecular Fingerprinting .............. 341<br />

19.5.2 Sample Preparation .............................................. 342<br />

19.5.3 Nucleic Acid Extraction......................................... 343<br />

19.5.4 PCR and Molecular Community Fingerprinting........ 344<br />

19.5.5 Group-Specific Molecular<br />

Community Fingerprinting ................................... 345<br />

19.5.6 Molecular Identification of Species and Genes .......... 348<br />

19.6 Integration of Detection Techniques .................................. 348<br />

19.6.1 Polyphasic Approach ............................................ 348<br />

19.7 Conclusions ................................................................... 349<br />

<strong>References</strong> ............................................................................. 350<br />

Subject Index 355


Contributors<br />

Anand, Richa<br />

Faculty of Land and Food Sciences, Faculty of Forestry, University of British<br />

Columbia, Vancouver, British Columbia, Canada V6T 1Z4; Current address:<br />

Department of Forest Mycology and Pathology, Swedish University of Agricultural<br />

Sciences (SLU), Box 7026, 750 07 Uppsala, Sweden<br />

Bacon, Charles W.<br />

Richard B. Russell Research Center, ARS, United States Department of<br />

Agriculture, Toxicology and Mycotoxin Research Unit, SAA, P.O. Box 5677,<br />

Athens, GA 30604, USA<br />

Bärlocher, Felix<br />

63BYorkStreet,DepartmentofBiology,MountAllisonUniversity,Sackville,<br />

New Brunswick, E4L 1G7, Canada<br />

Bayman, Paul<br />

Departamento de Biologia, Universidad de Puerto Rico – Rio Piedras, PO<br />

Box 23360, San Juan, PR 00931, USA<br />

Berg, Gabriele<br />

Graz University of Technology, Department of Environmental Biotechnology,<br />

Petersgasse 12, 8010 Graz, Austria<br />

Bloemberg, Guido V.<br />

Leiden University, Institute of Biology, Wassenaarseweg 64, 2333AL Leiden,<br />

The Netherlands<br />

Boyle, Christine<br />

Augustastraße 32, 02826 Görlitz, Germany<br />

Broughton, William J.<br />

Université de Genève, Laboratoire de Biologie Moléculaire des Plantes<br />

Supérieures, Sciences III, 30 Quai Ernest-Ansermet, 1211 Genève 4, Switzerland


XVI Contributors<br />

Brundrett, Mark C.<br />

School of Plant Biology, Faculty of Natural and Agricultural Sciences, The<br />

University of Western Australia, Crawley, WA 6009, Australia<br />

Cairney, John W.G.<br />

Centre for Plant and Food Sciences, Parramatta Campus, University of<br />

Western Sydney, Locked Bag 1797, Penrith South DC, NSW 1797, Australia<br />

Carvajal, Margarita M. Camacho<br />

Leiden University, Institute of Biology, Wassenaarseweg 64, 2333AL Leiden,<br />

The Netherlands<br />

Chanway, Chris<br />

Faculty of Land and Food Systems, Faculty of Forestry, University of British<br />

Columbia, Vancouver, British Columbia, Canada V6T 1Z4<br />

Currah, Randolph S.<br />

Department of Biological Sciences, University of Alberta, Edmonton, AB,<br />

T6G 2E9, Canada<br />

Deakin, William J.<br />

Université de Genève, Laboratoire de Biologie Moléculaire des Plantes<br />

Supérieures, Sciences III, 30 Quai Ernest-Ansermet, 1211 Genève 4, Switzerland<br />

Elsas, Jan D. van<br />

Groningen University, Department of Microbial Ecology, Biological Center,<br />

Kerklaan 30, 9751 NN, Haren, The Netherlands<br />

Girlanda, Mariangela<br />

Dipartimento di Biologia Vegetale and IPP - Torino, Viale PA Mattioli 25,<br />

10125 Torino, Italy<br />

Grünig, Christoph R.<br />

Swiss Federal Institute of Technology, Department of Environmental Sciences,<br />

Institute of Integrative Biology, Forest Pathology and Dendrology,<br />

8092 Zürich, Switzerland<br />

Hallmann, Johannes<br />

Biologische Bundesanstalt für Land- und Forstwirtschaft, Institut für Nematologie<br />

und Wirbeltierkunde, Toppheideweg 88, 48161 Münster, Germany


Contributors XVII<br />

Jansson, Hans-Börje<br />

Departamento de Ciencias del Mar y Biología Aplicada, Universidad de<br />

Alicante, Apdo. 99, 03080 Alicante, Spain<br />

Kloepper, Joseph W.<br />

Department of Entomology and Plant Pathology, Auburn University, Auburn,<br />

AL 36849, USA<br />

Lopez-Llorca, Luis V.<br />

Departamento de Ciencias del Mar y Biología Aplicada, Universidad de<br />

Alicante, Apdo. 99, 03080 Alicante, Spain<br />

Luppi, Anna Maria<br />

Dipartimento di Biologia Vegetale and IPP, Viale PA Mattioli 25, 10125<br />

Torino, Italy<br />

Otero, J. Tupac<br />

Universidad Nacional de Colombia-Palmira, Departamento de Ciencias<br />

Agrícolas, AA 237, Palmira, Valle del Cauca, Colombia<br />

Overbeek, Leo S. van<br />

Plant Research International B.V, Droevendaalsesteeg 1, 6708 PB, Wageningen,<br />

The Netherlands<br />

Paul, Leslie<br />

Faculty of Land and Food Sciences, Systems, University of British Columbia,<br />

Vancouver, British Columbia, Canada V6T 1Z4<br />

Perotto, Silvia<br />

Dipartimento di Biologia Vegetale and IPP - Torino, Viale PA Mattioli 25,<br />

10125 Torino, Italy<br />

Rice, Adrianne V.<br />

Northern Forestry Centre, Canadian Forest Service, Natural Resources<br />

Canada, 5320-122 St., Edmonton, AB, T6H 3S5 Canada<br />

Ryu, Choong-Min<br />

Plant Biology Division, The Samuel Roberts Noble Foundation, 2510 Sam<br />

Noble Parkway, Ardmore, OK 73401, USA


XVIII Contributors<br />

Saad, Maged M.<br />

Université de Genève, Laboratoire de Biologie Moléculaire des Plantes<br />

Supérieures, Sciences III, 30 Quai Ernest-Ansermet, 1211 Genève 4, Switzerland<br />

Salinas, Jesús<br />

Departamento de Ciencias del Mar y Biología Aplicada, Universidad de<br />

Alicante, Apdo. 99, 03080 Alicante, Spain<br />

Schulz, Barbara<br />

Institute of Microbiology, Technical University of Braunschweig, Spielmannstraße<br />

7, 38106 Braunschweig, Germany<br />

Sieber, Thomas N.<br />

Swiss Federal Institute of Technology, Department of Environmental Sciences,<br />

Institute of Integrative Biology, Forest Pathology and Dendrology,<br />

8092 Zürich, Switzerland<br />

Vicente, José Gaspar Maciá<br />

Departamento de Ciencias del Mar y Biología Aplicada, Universidad de<br />

Alicante, Apdo. 99, 03080 Alicante, Spain<br />

Vuurde, Jim van<br />

Plant Research International B.V, Droevendaalsesteeg 1, 6708 PB, Wageningen,<br />

The Netherlands<br />

Yates, Ida E.<br />

Richard B. Russell Research Center, ARS, United States Department of<br />

Agriculture, Toxicology and Mycotoxin Research Unit, Athens, GA 30604,<br />

USA


Abbreviations<br />

ACC 1-aminocyclopropane-1-carboxylate<br />

AHL acyl homoserine lactones<br />

AM arbuscular mycorrhiza<br />

AMF arbuscular mycorrhizal fungi<br />

ARA acetylene reduction activity<br />

AUDPC area under the disease progress curve<br />

BAC bacterial artificial chromosome<br />

BCA biocontrol agents<br />

BRD root border cells<br />

BrdU bromide oxyuridine<br />

Cfu colony forming units<br />

CLSM confocal laser scanning microscopy<br />

CMA corn meal agar<br />

CMV Cucumber mosaic virus<br />

CPS capsular polysaccharides<br />

CTAB cetyl trimethyl ammonium bromide<br />

DGGE denaturing gradient gel electrophoresis<br />

DIC differential interference microscopy<br />

DON deoxynivalenol<br />

DSE dark septate endophytes<br />

DSF dark septate fungi<br />

DSM dark sterile mycelia<br />

ECFP enhanced cyan fluorescent protein<br />

ECM ectomycorrhizae<br />

ECM extracellular material<br />

EGFP Enhanced GFP<br />

ELISA enzyme-linked immunosorbent assay<br />

EPS extra-cellular polysaccharides<br />

EYFP Enhanced Yellow Fluorescent Protein<br />

FISH fluorescence in situ hybridization<br />

GFP Green fluorescent protein<br />

GSP general secretory pathway<br />

GUS β-glucuronidase


XX Abbreviations<br />

IAA indole acetic acid<br />

ICR induced systemic resistance<br />

ISSR inter-simple sequence repeat<br />

ITS internal transcribed spacer<br />

ITS-RFLP internal transcribed spacer-restriction fragment length<br />

polymorphism<br />

LPS lipo-polysaccharides<br />

MLH multilocus haplotypes<br />

MRA Mycelium radicis atrovirens<br />

NDFA nitrogen derived from the atmosphere<br />

PAH polyaromatic hydrocarbon<br />

PAR photosynthetically active radiation<br />

PBM peribacteroid membrane<br />

PCR polymerase chain reaction<br />

PCR-RFLP polymerase chain reaction -restriction fragment length<br />

polymorphism<br />

PDA potato dextrose agar<br />

PGPR plant growth promoting rhizobacteria<br />

PR pathogenesis-related<br />

PVP polyvinylpyrrolidone<br />

RAPD random amplified polymorphic DNA<br />

RDNA 16S rRNA gene<br />

RFLP restriction fragment length polymorphism<br />

RFP red fluorescent protein<br />

RGR relative growth rate<br />

SAR systemic acquired resistance<br />

SEM Scanning electron microscopy<br />

SPB sterile phosphate buffer<br />

SPS Surface polysaccharides<br />

SSCP single strand conformational polymorphism<br />

TGGE temperature gradient gel electrophoresis<br />

TNV tobacco necrosis virus<br />

ToMoV Tomato mottle virus<br />

TRFLP terminal restriction fragment length polymorphism<br />

TSA tryptic soya agar<br />

VAM vesicular arbuscular mycorrhiza<br />

VOC volatile organic compound<br />

VWT variable white taxon<br />

WA Western Australia


1<br />

What are Endophytes?<br />

Barbara Schulz, Christine Boyle<br />

1.1<br />

Introduction and Definitions<br />

Taken literally, the word endophyte means “in the plant” (endon Gr. =<br />

within, phyton = plant). The usage of this term is as broad as its literal<br />

definition and spectrum of potential hosts and inhabitants, e.g. bacteria<br />

(Kobayashi and Palumbo 2000), fungi (Stone et al. 2000), plants (Marler<br />

et al. 1999) and insects in plants (Feller 1995), but also for algae within<br />

algae (Peters 1991). Any organ of the host can be colonised. Equally variable<br />

is the usage of the term “endophyte” for variable life history strategies<br />

of the symbiosis, ranging from facultatively saprobic to parasitic to<br />

exploitive to mutualistic. The term endophyte is, for example, used for<br />

pathogenic endophytic algae (Bouarab et al. 1999), parasitic endophytic<br />

plants (Marler et al. 1999), mutualistic endophytic bacteria (Chanway 1996;<br />

Adhikari et al. 2001; Bai et al. 2002) and fungi (Carroll 1988; Jumpponen<br />

2001; Sieber 2002; Schulz and Boyle 2005), and pathogenic bacteria and<br />

fungi in latent developmental phases (Sinclair and Cerkauskas 1996), but<br />

also for microorganisms in commensalistic symbioses (Sturz and Nowak<br />

2000).<br />

Some authors also designate the interactions of mycorrhizal fungi with<br />

the roots of their hosts as being endophytic (reviewed by Sieber 2002).<br />

However, we concur with Brundrett (2004; see Chap. 16 by Brundrett),<br />

who distinguishes mycorrhizal from endophytic interactions; the former<br />

having synchronised plant-fungus development and nutrient transfer at<br />

specialised interfaces. Nevertheless, as we will see in this book, distinctions<br />

between mycorrhizal and non-mycorrhizal fungi are not always clear-cut<br />

[see Chaps. 9 (Bayman and Otero), 12 (Girlanda et al.), 13 (Rice and Currah),<br />

14 (Cairney), and 15 (Schulz)]. Not only can mycorrhizal fungi become<br />

pathogenic, but, for example, dark septate endophytes (DSE) can assume<br />

mycorrhizal functions [Jumpponen and Trappe 1998; see Chaps. 7 (Sieber<br />

Barbara Schulz: Technical University of Braunschweig, Institute of Microbiology, Spielmannstraße<br />

7, 38106 Braunschweig, Germany, E-mail: b.schulz@tu-bs.de<br />

Christine Boyle: Augustastraße 32, 02826 Görlitz, Germany<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


2 B. Schulz, C. Boyle<br />

and Grünig), and 15 (Schulz)]. In addition, there are also cases in which<br />

fungal root endophytes seem to be saprobes, e.g. Oidiodendron maius (see<br />

Chap.13byRiceandCurrah)andPhialocephala fortinii (Jumpponen and<br />

Trappe 1998; Jumpponen et al. 1998; see Chap. 15 by Schulz).<br />

Although there are diverse uses for the word endophyte, “endophytes”<br />

are most commonly defined as those organisms whose “...infections are<br />

inconspicuous, the infected host tissues are at least transiently symptomless,<br />

and the microbial colonisation can be demonstrated to be internal...”<br />

(Stone et al. 2000). Although these authors used this definition to describe<br />

fungal endophytes, it is equally applicable to bacterial endophytes.<br />

It is important to remember that the definition describes a momentary<br />

status. Thus it includes an assemblage of microorganisms with different<br />

life history strategies: those that grow saprophytically on dead or senescing<br />

tissues following an endophytic growth phase (Stone 1987; see Chap. 8 by<br />

Bacon and Yates), avirulent microorganisms as well as latent pathogens and<br />

virulent pathogens in the early stages of infection (Sinclair and Cerkauskas<br />

1996; Kobayashi and Palumbo 2000). Unfortunately, taken literally, it can<br />

include all pathogens at some stage of their development. Since the plant<br />

host responds to at least some infections with mechanical defence reactions<br />

(Narisawa et al. 2004; see Chap. 15 by Schulz), there is merit to Petrini’s<br />

additional characterisation of endophytic interactions as not “causing apparent<br />

harm” (Petrini 1991), which presumably refers to an absence of<br />

macroscopically visible symptoms. Aware of the determinative discrepancies,<br />

we will nevertheless use the term “endophyte” to describe those<br />

bacteria and fungi that can be detected at a particular moment within the<br />

tissues of apparently healthy plant hosts (Schulz and Boyle 2005).<br />

1.2<br />

Colonisation<br />

In spite of the fact that bacteria are prokaryotes and fungi are eukaryotes,<br />

they share many attributes of their associations with plant hosts, e.g. both<br />

colonise root tissues inter- and intra-cellularly, and often systemically (Table<br />

1.1). They do, however, differ somewhat in their modes of colonisation.<br />

Bacteria primarily colonise intercellularly (Hinton and Bacon 1995; Hallmann<br />

et al. 1997), though they have also been found intracellularly, e.g.<br />

Azoarcus spp. (Hurek et al. 1994). They are frequently found in the vascular<br />

tissues of host plants (Kobayashi and Palumbo 2000), which is advantageous<br />

for distribution, whereas asymptomatic colonisation by fungi may<br />

be inter- and intra-cellular throughout the root. Although DSE sometimes<br />

colonise the vascular cylinder in asymptomatic interactions (Barrow 2003),<br />

such colonisation is frequently associated with pathogenicity (Bacon and<br />

Hinton 1996; Schulz and Boyle 2005).


1WhatareEndophytes? 3<br />

Table 1.1. Characteristics of the interactions of bacterial vs. fungal endophytes with plant roots (see also all other book chapters)<br />

Criteria Bacteria Fungi<br />

Host spectrum Broad, depends on habitat, host, season Broad, depends on habitat, host, season<br />

Mode and site of infection Passive through wounds and other tissue openings Active: through stomata, cell wall or wounds<br />

or active with enzymes or vectors, e.g. insects<br />

Nutritional source<br />

Host exudates, dead cortex cells, plant debris Storage material in the spores,<br />

during first stage infection<br />

dead cortex cells, plant debris, host exudates<br />

Nutritional source<br />

Components of the symplast and apoplast Components of the symplast and apoplast<br />

during colonisation, i.e.<br />

during a “steady state” status<br />

Growth in root Inter- and/or intra-cellular, slow, low colonisation densities Inter- and/or intracellular, often extensive<br />

Growth from roots into the shoot Yes Sometimes<br />

Systemic growth in roots Possible Possible<br />

Tissue colonised Primarily intercellular, also vascular tissue Usually not within vascular tissue<br />

Specialised structures<br />

Nodules, glands Sometimes<br />

for nutrient access<br />

Physiological status Only little data available Balanced antagonisms, active interaction<br />

Outcome of the interaction Commensalism, mutualism or latent pathogenicity Commensalism, mutualism or latent pathogenicity<br />

Benefits<br />

A reliable supply of nutrients and protection from A reliable supply of nutrients and protection from<br />

for the microbial symbiont environmental stresses, passive transfer and spread environmental stresses, advantages for reproduction<br />

between hosts via vectors, e.g. insects<br />

and colonisation at host senescence<br />

Potential benefits<br />

Induced resistance, improved growth (N-fixation, Induced resistance, improved growth<br />

for the plant symbiont<br />

phytohormones), synthesis of metabolites antagonistic to (phytohormones, improved access to minerals and<br />

plant pathogens and parasites<br />

nutrients), synthesis of metabolites antagonistic to<br />

predators and antagonists<br />

Reproduction Usually passive transfer and spread between hosts via Active and passive following host senescence,<br />

vectors, e.g. insects, but also active, e.g. Pseudomonads sometimes with vectors


4 B. Schulz, C. Boyle<br />

The assemblages of fungi that colonise plant roots are diverse (Vandenkoornhuyse<br />

et al. 2002). In contrast to endophytic growth in the aboveground<br />

plant organs, endophytic growth of fungi within the roots has<br />

frequently been found to be extensive (Stone et al. 2000; Schulz and Boyle<br />

2005; see Chap. 11 by Lopez-Llorca et al.). Root colonisation can be both<br />

inter- and intra-cellular, the hyphae often forming intracellular coils, e.g.<br />

DSE (Jumpponen and Trappe 1998; Stone et al. 2000; Sieber 2002), the basidiomycete<br />

Piriformospora indica (Varma et al. 2000), or Oidiodendron<br />

maius (see Chap. 13 by Rice and Currah) and Heteroconium chaetospira<br />

(Usuki and Narisawa 2005), which can even form characteristic ericoid<br />

mycorrhizal infection units (see Chap. 14 by Cairney). DSE may also form<br />

ectendomycorrhiza (Lubuglio and Wilcox 1988) and ectomycorrhizal-like<br />

structures (Wilcox and Wang 1987; Fernando and Currah 1996; Kaldorf et<br />

al. 2004; see Chap. 15 by Schulz).<br />

Many orchid roots are systemically and mycoheterotrophically colonised<br />

by fungi of the genus Rhizoctonia (Ma et al. 2003; see Chap. 16 by Brundrett)<br />

and Leptodontidium (Bidartondo et al. 2004). In some cases, e.g. Fusarium<br />

verticillioides (= F. moniliforme), colonisation by an avirulent strain was<br />

found to be systemic and intercellular, whereas pathogenic strains also<br />

colonised intracellularly (Bacon and Hinton 1996). Latent pathogens, e.g.<br />

Cryptosporiopsis sp. (Kehr 1992; Verkley 1999) may occasionally penetrate<br />

the vascular bundles (Schulz and Boyle 2005).<br />

Bacteria usually invade the roots passively, e.g. at open sites on roots such<br />

as lateral root emergence or wounds (Kobayashi and Palumbo 2000), even<br />

achieving systemic colonisation from a single site of entry (Hallmann et al.<br />

1997). Although colonisation densities of nonpathogenic endophytic bacteria<br />

are rarely as high as those of pathogenic bacteria, they are highest in the<br />

root tissue; perhaps because this is the primary site of infection (Kobayashi<br />

and Palumbo 2000; Hallmann et al. 1997; see Chap. 2 by Hallmann and<br />

Berg).<br />

1.3<br />

Assemblages and Adaptation<br />

Both fungal and bacterial endophytes have been isolated from the roots of<br />

almost all hosts studied to date [Petrini 1991; Stone et al. 2000; Kobayashi<br />

and Palumbo 2000; Sieber 2002; see Chaps. 2 (Hallmann and Berg),<br />

3 (Kloepper and Ryu), and 7 (Sieber and Grünig)]. The assemblages of<br />

endophytes that colonise a particular host vary both with habitat and host,<br />

some even being adapted to very specialised habitats, e.g. the aquatic fungi<br />

that colonise submerged roots (see Chap. 10 by Bärlocher). Recent molecular<br />

methods enable better analyses of the geographical distribution of given


1WhatareEndophytes? 5<br />

groups of microorganisms, for example that of the DSE [Jumpponen 1999,<br />

see Chaps. 7 (Sieber and Grünig), 12 (Girlanda et al.), and 15 (Cairney)].<br />

Both diversity and colonisation density frequently increase during the<br />

course of the vegetation period (Smalla et al. 2001), since horizontal transmission<br />

predominates (Carroll 1988, 1995; Petrini 1991; Guske et al. 1996;<br />

Hallmann et al. 1997; Arnold and Herre 2003, see Chap. 2 by Hallmann and<br />

Berg). Particularly asexual sporulation increases in autumn at the end of<br />

the vegetation period.<br />

Communities of endophytes inhabiting a particular host may be ubiquitous,<br />

or have what is frequently referred to as host specificity (e.g. Carroll<br />

1988; Petrini 1996; Stone et al. 2000; Berg et al. 2002; Cohen 2004). We concur<br />

with Carroll (1999) and Zhou and Hyde (2001) that the term “specificity”<br />

should be reserved for organisms that will only grow in one host (Schulz<br />

and Boyle 2005). If this is not the case, this phenomenon could be termed<br />

host preference (Carroll 1999) or host-exclusivity (Zhou and Hyde 2001).<br />

Whether the interaction represents specificity, preference or exclusivity,<br />

an adaptation of host and endophyte to one another has occurred. The<br />

adaptation may not only be to a particular host, but to endophytic growth<br />

in one plant organ, e.g. in the roots in contrast to the shoots [Petrini 1991;<br />

Hallmann et al. 1997; Sieber 2002; Schulz and Boyle 2005; see Chaps. 2 (Hallmann<br />

and Berg), and 7 (Sieber and Grünig)].<br />

It is often extremely difficult to know whether or not a particular fungus<br />

or bacterium that has been detected in healthy plant tissue has actually<br />

been growing within the host tissue or has been incidentally isolated, i.e.<br />

is normally found on other substrates. As reviewed by Schulz and Boyle<br />

(2005) and in Chap. 17 by Hallmann et al., there are four methods presently<br />

in use for detecting and identifying fungi and bacteria in plant tissue:<br />

(1) histological observation (see Chap. 6 by Anand et al.), most recently<br />

in combination with molecular methods (see Chap. 18 by Bloemberg and<br />

Carvajal), (2) surface sterilisation of the host tissue and isolation of the<br />

emerging fungi on appropriate growth media, (3) detection by specific<br />

chemistry, e.g. immunological methods (see Chap. 18 by Bloemberg and<br />

Carvajal), or (4) by direct amplification of fungal DNA from colonised<br />

plant tissues [Vandenkoornhuyse et al. 2002; see Chaps. 17 (Hallmann et<br />

al.), and 19 (van Overbeek et al.)], having first ascertained that there are<br />

no fungal residues on the plant surface (Arnold et al. 2006). Methods for<br />

quantification are reviewed by Sieber (2002), Schulz and Boyle (2005) and<br />

in Chap. 17 (Hallmann et al.).


6 B. Schulz, C. Boyle<br />

1.4<br />

Life History Strategies<br />

Organismsdetectedatanyonemomentinasymptomaticplanttissueand<br />

arbitrarily named “endophytes” include microorganisms with different<br />

life history strategies. Endophytes represent, both as individuals and collectively,<br />

a continuum of mostly variable associations: mutualism, commensalism,<br />

latent pathogenicity, and exploitation. The phenotypes of the<br />

interactions are often plastic, depending on the genetic dispositions of the<br />

two partners, their developmental stage and nutritional status, but also on<br />

environmental factors (see Chap. 12 by Girlanda et al.). The role of genetic<br />

disposition was demonstrated by Freeman and Rodriguez (1993): a single<br />

mutation resulted in loss of a virulence factor, transforming a pathogenic<br />

fungus, Colletotrichum magna, intoanendophyte.Similarly,avirulence<br />

genes and the machinery of pathogenicity may be lacking or suppressed in<br />

bacterial endophytes (Kobayashi and Palumbo 2000).<br />

Just as fungi have been found to develop ectomycorrhiza in one host<br />

and what appear to be ericoid mycorrhiza in another host (Villarreal-<br />

Ruiz et al. 2004), a mycorrhizal fungus can grow endophytically in the<br />

roots of a non-host (see Chap. 12 by Girlanda et al.). The importance<br />

of a particular combination of host and microorganism as well as their<br />

reciprocal influences also becomes apparent when a fungal or bacterial<br />

pathogen is inoculated into a non-host and is no longer virulent, colonising<br />

as an asymptomatic endophyte (Carroll 1999; Kobayashi and Palumbo 2000;<br />

Schulz and Boyle 2005) The influence of the host plant in determining the<br />

mycorrhizal, endophytic or even pathogenic character of a DSE association<br />

is likely to be a prime factor. In plant communities, the multiple mutualistic<br />

potential of these fungi, establishing hyphal links or inoculum reservoirs,<br />

may favour inter-plant interactions (see Chap. 12 by Girlanda et al.).<br />

Interactions are frequently complex, involving more than two partners.<br />

Endophytic bacteria and fungi may interact not only with the plant host,<br />

but also with other organisms, including mycorrhizal fungi (see Chap. 9<br />

by Bayman and Otero) and metazoa. For example, nematophagous fungi,<br />

which are ubiquitous organisms in soils, not only can switch from a saprophytic<br />

to a parasitic stage to kill and digest living nematodes, but can also<br />

grow endophytically in plant roots (see Chap. 11 by Lopez-Llorca et al.).<br />

Mutualistic interactions involving fungi and bacteria that endophytically<br />

colonise plant roots benefit the microbial partner with a reliable supply of<br />

nutrients as well as protection from environmental stresses. As reported<br />

in this book, benefits for the host plant may include improved growth [see<br />

Chaps. 6 (Anand et al.), 13 (Rice and Currah), 15 (Schulz), and 19 (van<br />

Overbeek et al.)], induced resistance [see Chaps. 3 (Kloepper and Ryu),<br />

4 (Berg and Hallmann), 6 (Anand et al.), and 15 (Schulz)], biocontrol of


1WhatareEndophytes? 7<br />

plant parasitic nematodes (see Chap. 11 by Lopez-Llorca et al.) and of fungi<br />

in agriculture [see Chaps. 3 (Kloepper and Ryu), 4 (Berg and Hallmann),<br />

15 (Schulz)] and forestry (see Chap. 6 by Anand et al.), as well as microbial<br />

synthesis of metabolites antagonistic to predators [Schulz et al. 2002; Schulz<br />

and Boyle 2005; see Chaps. 6 (Anand et al.), 8 (Bacon and Yates), 15 (Schulz),<br />

and 19 (van Overbeek et al.)] When synthesized in agricultural crops in<br />

situ, mycotoxins synthesised by endophytes, e.g. Fusarium verticillioides in<br />

maize, are potentially problematic for human consumption of these crops<br />

(see Chap. 8 by Bacon and Yates).<br />

Factors responsible for improving plant growth are the microbial synthesis<br />

of phytohormones [Tudzynski 1997; Tudzynski and Sharon 2002;<br />

Kobayashi and Palumbo 2000; see Chaps. 6 (Anand et al.) and 15 (Schulz)],<br />

access to minerals and/or other nutrients from the soil (Caldwell et al. 2000;<br />

Barrow 2003; see Chap. 13 by Rice and Currah), bacterial fixation of atmospheric<br />

nitrogen, which has been demonstrated not only for the noduleforming<br />

members of the Rhizobiaceae, but also for non-nodule-forming<br />

bacteria, e.g. Acetobactor and Azoarcus (Reinhold-Hurek and Hurek 1998;<br />

see Chap. 5 by Saad et al.). In the associations of nitrogen-fixing rhizobia<br />

with legumes, some of the same signalling molecules are involved as in<br />

the interactions of mycorrhizal fungi with their hosts, e.g. flavonoids and<br />

nod-factors (Lapopin and Franken 2000; Martin et al. 2001; Mirabella et<br />

al. 2002; Imaizumi-Anraku et al. 2005; see Chap. 5 by Saad et al.). And as<br />

has recently been shown, plastid membrane proteins involved in the first<br />

signalling interactions are crucial for the entry of both symbionts into the<br />

host roots (Imaizumi-Anraku et al. 2005).<br />

1.5<br />

Balanced Antagonism<br />

According to Heath (1997), only a few fungi are actually capable of causing<br />

disease in any one plant, since they must first cross several barriers and<br />

overcome other plant defences. This must also be true for bacteria. Thus,<br />

one question has motivated many investigations: how does the endophyte<br />

manage to exist, and often to grow, within its host without causing visible<br />

disease symptoms? We have proposed a working hypothesis based on observations<br />

from the interactions studied thus far (Schulz et al. 1999; Schulz<br />

and Boyle 2005). Asymptomatic colonisation is a balance of antagonisms<br />

between host and endophyte (Fig. 1.1). Endophytes and pathogens both<br />

possess many of the same virulence factors: the endophytes studied thus far<br />

produced the exoenzymes necessary to infect and colonise the host (Sieber<br />

et al. 1991; Petrini et al. 1992; Ahlich-Schlegel 1997; Boyle et al. 2001; Lumyong<br />

et al. 2002), even though only some of these endophytes are presumably


8 B. Schulz, C. Boyle<br />

Fig.1.1. Hypothesis: a balance of antagonisms between endophytic virulence and plant<br />

defence response results in asymptomatic colonisation (reproduced with permission from<br />

Schulz and Boyle 2005)<br />

latent pathogens. The majority can produce phytotoxic metabolites (Schulz<br />

et al. 2002; Schulz and Boyle 2005). The host can respond with the same defence<br />

reactions as to a pathogen, i.e. with preformed and induced defence<br />

metabolites [Yates et al. 1997; Schulz et al. 1999; Mucciarelli et al. 2003; see<br />

Chaps. 3 (Kloepper and Ryu), and 11 (Lopez-Llorca et al.)], and general<br />

defence responses (Narisawa et al. 2004; Schulz and Boyle 2005). As long as<br />

fungal virulence and plant defence are balanced, the interaction remains<br />

asymptomatic. In all of these interactions we are referring to a momentary<br />

status, an often fragile balance of antagonisms.<br />

If the host-pathogen interaction becomes imbalanced, either disease results<br />

or the fungus is killed. In some cases, the virulence of weak pathogens<br />

such as Pezicula spp. (Kehr 1992) is sufficient for disease development only<br />

when the host is stressed or senescent. Whether the interaction is balanced<br />

or imbalanced depends on the general status of the partners, the virulence<br />

of the fungus, and the defences of the host – both virulence and defence<br />

being variable and influenced by environmental factors, nutritional status<br />

and developmental stages of the partners. Although this hypothesis has<br />

been developed to explain the interactions of fungal endophytes with their


1WhatareEndophytes? 9<br />

hosts, further studies may well provide evidence that it is also applicable to<br />

endophytic bacteria.<br />

Balanced antagonistic interactions are plastic in expression, depending<br />

on the momentary status of host and endophyte, but also on biotic and abiotic<br />

environmental factors and on the tolerance of each of the partners to<br />

these factors. In particular, many endophytes seem to be masters of phenotypic<br />

plasticity: infecting as a pathogen, colonising cryptically, and finally<br />

sporulating as a pathogen or saprophyte. This necessitates a balance with<br />

the potential for variability, which means that these endophytic interactions<br />

are creative, having the potential for evolutionary development – the<br />

symbioses can evolve both in the direction of more highly specialised mutualisms<br />

and in the direction of more highly specialised parasitisms and<br />

exploitation. Indeed, there is evidence that mycorrhizal fungi may have<br />

evolved from the endophytic activity of saprophytic fungi (see Chap. 16<br />

by Brundrett), but also that plastids that have evolved from endosymbiotic<br />

bacteria facilitate further symbioses with other bacterial and fungal<br />

symbionts (Imaizumi-Anraku et al. 2005).<br />

1.6<br />

Conclusions<br />

The usage of the term “endophyte” is as broad as its literal definition<br />

and spectrum of potential hosts and inhabitants. The most common usage<br />

of the term “endophyte” for organisms whose infections are internal and<br />

inconspicuous, and in which the infected host tissues are at least transiently<br />

symptomless, is equally applicable to bacterial prokaryotes and fungal<br />

eukaryotes.<br />

Endophytes include an assemblage of microorganisms with different life<br />

history strategies: those that, following an endophytic growth phase, grow<br />

saprophytically on dead or senescing tissue, avirulent microorganisms,<br />

incidentals, but also latent pathogens and virulent pathogens at early stages<br />

of infection. These parasitic interactions may vary from mutualistic to<br />

commensalistic to latently pathogenic and exploitive. Phenotypes of the<br />

interactions are often plastic, depending on the genetic dispositions of the<br />

two partners, their developmental stage and nutritional status, but also on<br />

environmental factors.<br />

We have proposed a working hypothesis based on observations from the<br />

interactions studied thus far to explain asymptomatic microbial colonisation<br />

as a balance of antagonisms between host and endophyte (Fig. 1.1;<br />

Schulz and Boyle 2005). This often fragile balance of antagonism is a momentary<br />

status and depends on the general status of the partners, the<br />

virulence of the fungus and defences of the host, environmental fac-


10 B. Schulz, C. Boyle<br />

tors, nutritional status, as well as the developmental stages of the partners.<br />

<strong>References</strong><br />

Adhikari TG, Joseph CM, Yang G, Philips DA, Nelson LM (2001) Evaluation of bacteria<br />

isolated from rice for plant growth promotion and biological control of seedling disease<br />

of rice. Can J Microbiol 47:916–924<br />

Ahlich-Schlegel K (1997) Vorkommen und Charakterisierung von dunklen, septierten<br />

Hyphomyceten (DSH) in Gehölzwurzeln. Dissertation, Eidgenössische Technische<br />

Hochschule, Zürich, Switzerland<br />

Arnold AE, Herre EA (2003) Canopy cover and leaf age affect colonisation by tropical fungal<br />

endophytes: ecological pattern and process in Theobroma cacao (Malvaceae) Mycologia<br />

95:388–398<br />

Arnold AE, Henk DA, Eells RL, Lutzoni F, Vilgalys R (2006) Diversity and phylogenetic affinities<br />

of foliar fungal endophytes in loblolly pine inferred by culturing and environmental<br />

PCR. Mycologia (in press)<br />

Bacon CW, Hinton NS (1996) Symptomless endophytic colonisation of maize by Fusarium<br />

moniliforme. Can J Bot 74:1195–1202<br />

Bai Y, Aoust F, Smith D, Driscoll B (2002) Isolation of plant-growth-promoting Bacillus<br />

strains from soybean root nodules. Can J Microbiol 48:230–238<br />

Barrow JR (2003) Atypical morphology of dark septate fungal root endophytes of Bouteloua<br />

in arid southwestern USA rangelands. Mycorrhiza 13:239–247<br />

Berg G, Roskot N, Steidle A, Eberl L, Zock A, Smalla K (2002) Plant-dependent genotypic and<br />

phenotypic diversity of antagonistic rhizobacteria isolated from different Verticillium<br />

host plants. Appl Environ Microbiol 68:3328–3338<br />

Bidartondo ML, Burghardt B, Gebauer G, Bruns TD, Read DJ (2004) Changing partners in<br />

the dark: isotopic and molecular evidence of ectomycorrhizal liaisons between forest<br />

orchids and trees. Proc R Soc London B 271:1799–1806<br />

Bouarab K, Potin P, Correa J, Kloareg B (1999) Sulfated oligosaccharides mediate the interaction<br />

between a marine red alga and its green algal pathogenic endophyte. Plant Cell<br />

11:1635–1650<br />

Boyle C, Götz M, Dammann-Tugend U, Schulz B (2001) Endophyte–host interactions III.<br />

Local vs. systemic colonisation. Symbiosis 31:259–281<br />

Brundrett MC (2004) Diversity and classification of mycorrhizal associations. Biol Rev<br />

79:473–495<br />

Caldwell BA, Jumpponen A, Trappe JM (2000) Utilization of major detrital substrates by<br />

dark-septate, root endophytes. Mycologia 92:230–232<br />

Carroll GC (1988) Fungal endophytes in stems and leaves: from latent pathogen to mutualistic<br />

symbiont. Ecology 69:2–9<br />

Carroll GC (1995) Forest endophytes: pattern and process. Can J Bot 73:1316–1324<br />

Carroll GC (1999) The foraging Ascomycete. XVI International Botanical Congress,<br />

St Louis, MN<br />

Chanway CP (1996) Endophytes: they’re not just fungi! Can J Bot 74:321–322<br />

Cohen SD (2004) Endophytic-host selectivity of Discula umbrinella on Quercus alba and<br />

Quercus rubra characterized by infection, pathogenicity and mycelial compatibility. Eur<br />

J Plant Pathol 110:713–721<br />

Feller IC (1995) Effects of nutrient enrichment on growth and herbivory of dwarf red<br />

mangrove (Rhizophora mangle). Ecol Monogr 65:477–505


1WhatareEndophytes? 11<br />

Fernando AA, Currah RS (1996) A comparative study of the effects of the root endophytes<br />

Leptodontidium and Phialocephala fortinii (Fungi imperfecti) on the growth of some<br />

subalpine plants in culture. Can J Bot 74:1071–1078<br />

Freeman S, Rodriguez RJ (1993) Genetic conversion of a fungal plant pathogen to a nonpathogenic,<br />

endophytic mutualist. Science 260:75–78<br />

Guske S, Boyle C, Schulz B (1996) New aspects concerning biological control of Cirsium<br />

arvense. IOBC/wprs Bulletin 19:281–290<br />

Hallmann J, Quadt-Hallmann A, Mahaffee WF, Kloepper JW (1997) Bacterial endophytes<br />

in agricultural crops. Can J Microbiol 43:895–914<br />

Heath MC (1997) Evolution of plant resistance and susceptibility to fungal parasites. In:<br />

Carroll GC, Tudzynski P (eds) The mycota V Part B Plant relationships. Springer, Berlin<br />

Heidelberg New York, pp 257–276<br />

Hinton DM, Bacon CW (1995) Enterobacter cloacae is an endophytic symbiont of corn.<br />

Mycopathologia 129:117–125<br />

Hurek T, Reinhold-Hurek B, Van Montagu M, Kellenberger E (1994) Root colonization and<br />

systemic spreading of Azoarcus sp. strain BH72 in grasses. J Bacteriol 176:1913–1923<br />

Imaizumi-Anraku H, Takeda N, Charpentier M, Perry J, Miwa H, Umehara Y, Kouchi H,<br />

Murakami Y, Mulder L, Vickers K, Pike J, Downie JA, Wang T, Sato S, Asamizu EE,<br />

Tabata S, Yoshikawa M, Murooka Y, Wu G, Kawaguchi M, Kawasaki S, Parniske M,<br />

Hayashi M (2005) Plastid proteins crucial for symbiotic fungal and bacterial entry into<br />

plant roots. Nature 433:527–531<br />

Jumpponen A (1999) Spatial distribution of discrete RAPD phenotypes of a root endophytic<br />

fungus, Phialocephala fortinii, at a primary successional site on a glacier forefront. New<br />

Phytol 141:333–344<br />

Jumpponen A (2001) Dark septate endophytes – are they mycorrhizal? Mycorrhiza<br />

11:207–211<br />

Jumpponen A, Trappe JM (1998) Performance of Pinus contorta inoculated with two strains<br />

of root endophytic fungus, Phialocephala fortinii: effects of synthesis system and glucose<br />

concentration. Can J Bot 76:1205–1213<br />

Jumpponen A, Mattson KG, Trappe JM (1998) Mycorrhizal functioning of Phialocephala<br />

fortinii with Pinus contorta on glacier forefront soil: interactions with soil nitrogen and<br />

organic matter. Mycorrhiza 7:261–265<br />

Kaldorf M, Renker C, Fladung M, Buscot F (2004) Characterization and spatial distribution of<br />

ectomycorrhizas colonizing aspen clones released in an experimental field. Mycorrhiza<br />

14:295–306<br />

Kehr RD (1992) Pezicula canker of Quercus rubra L, caused by Pezicula cinnamomea (DC)<br />

Sacc II Morphology and biology of the causal agent. Eur J For Pathol 22:29–40<br />

Kobayashi DY, Palumbo JD (2000) Bacterial endophytes and their effects on plants and uses<br />

in agriculture. In: Bacon CW, White JF (eds) Microbial endophytes. Dekker, New York,<br />

pp 199–236<br />

Lapopin L, Franken P (2000) Modification of plant gene expression. In: Kapulnik Y,<br />

Douds DD (eds) Arbuscular mycorrhizas: physiology and function. Kluwer, Dordrecht,<br />

pp 69–84<br />

Lubuglio KF, Wilcox HE (1988) Growth and survival of ectomycorrhizal and ectoendomycorrhizal<br />

seedlings of Pinus resinosa on iron tailings. Can J Bot 66:55–60<br />

Lumyong S, Lumyong P, McKenzie EHC, Hyde K (2002) Enzymatic activity of endophytic<br />

fungi of six native seedling species from Doi Suthep-Pui Nathional Park, Thailand. Can<br />

J Microbiol 48:1109–1112<br />

Ma M, Tan TK, Wong SM (2003) Identification and molecular phylogeny of Epulorhiza<br />

isolates from tropical orchids. Mycol Res 107:1041–1049


12 B. Schulz, C. Boyle<br />

Marler M, Pedersen D, Mitchell OT, Callaway RM (1999) A polymerase chain reaction<br />

method for detecting dwarf mistletoe infection in Douglas fir and western larch. Can J<br />

For Res 29:1317–1321<br />

Martin F, Duplessis S, Ditengou F, Lagrange H, Voiblet C, Lapeyrie F (2001) Developmental<br />

cross talking in the ectomycorrhizal symbiosis: signals and communication genes. New<br />

Phytol 151:145–154<br />

Mirabella R, Franssen H, Bisseling T (2002) LCO signalling in the interaction between<br />

rhizobia and legumes In: Scheel D, Wasternack C (eds) Plant signal transduction. Oxford<br />

University Press, Oxford, New York, pp 250–271<br />

Mucciarelli M, Scannerini S, Bertea C, Maffei M (2003) In vitro and in vivo peppermint<br />

(Mentha piperita) growth promotion by nonmycorrhizal fungal colonisation. New Phytol<br />

158:579–591<br />

Narisawa K, Usuki F, Hashiba T (2004) Control of Verticillium Yellows in Chinese cabbage<br />

by the dark septate endophytic fungus LtVB3. Phytopathology 94:412–418<br />

Peters AF (1991) Field and culture studies of Streblonema - Macrocystis new species Ectocarpales<br />

Phaeophyceae from Chile, a sexual endophyte of giant kelp. Phycologia<br />

30:365–377<br />

Petrini O (1991) Fungal endophytes of tree leaves. In: Andrews J, Hirano S (eds) Microbial<br />

ecology of leaves. Springer, New York Berlin Heidelberg, pp 179–197<br />

Petrini O (1996) Ecological and physiological aspects of host specificity in endophytic fungi<br />

In: Redlin SC, Carris LM (eds) Endophytic fungi in grasses and woody plants. APS,<br />

St Paul, MN, pp 87–100<br />

Petrini O, Sieber T, Toti L, Viret O (1992) Ecology, metabolite production, and substrate<br />

utilization in endophytic fungi. Nat Toxins 1:185–196<br />

Reinhold-Hurek B, Hurek T (1998) Life in grasses: diazotrophic endophytes. Trends Microbiol<br />

6:139–144<br />

Schulz B, Boyle C (2005) The endophytic continuum. Mycol Res 109:661–687<br />

Schulz B, Römmert A-K, Dammann U, Aust H-J, Strack D (1999) The endophyte-host<br />

interaction: a balanced antagonism. Mycol Res 103:1275–1283<br />

Schulz B, Boyle C, Draeger S, Römmert A-K, Krohn K (2002) Endophytic fungi: a source of<br />

biologically active secondary metabolites. Mycol Res 106:996–1004<br />

Sieber TN (2002) Fungal root endophytes In: Waisel Y, Eshel A, Kafkafi U (eds) The hidden<br />

half. Dekker, New York, pp 887–917<br />

Sieber TN, Sieber-Canavesi F, Petrini O, Ekramoddoullah AK, Dorworth CE (1991) Characterization<br />

of Canadian and European Melanconium from some Alnus by morphological<br />

cultural and biochemical studies. Can J Bot 69:2170–2176<br />

Sinclair JB, Cerkauskas RF (1996) Latent infection vs. endophytic colonisation by fungi. In:<br />

Redlin SC, Carris LM (eds) Endophytic fungi in grasses and woody plants. APS, St Paul,<br />

MN, pp 3–30<br />

Smalla K, Wieland G, Buchner A, Zock A, Parzy J, Roskot N, Heuer H, Berg G (2001) Bulk<br />

and rhizosphere soil bacterial communities studied by denaturing gradient gel electrophoresis:<br />

plant dependent enrichment and seasonal shifts. Appl Environ Microbiol<br />

67:4742–4751<br />

Stone JK (1987) Initiation and development of latent infections by Rhabdocline parkeri on<br />

Douglas-fir. Can J Bot 65:2614–2621<br />

Stone JK, Bacon CW, White JF (2000) An overview of endophytic microbes: endophytism<br />

defined. In: Bacon CW, White JF (eds) Microbial endophytes. Dekker, New York,<br />

pp 3–30<br />

Sturz AV, Nowak J (2000) Endophytic communities of rhizobacteria and the strategies<br />

required to create yield enhancing associations with crops. Appl Soil Ecol 15:183–190


1WhatareEndophytes? 13<br />

Tudzynski B (1997) Fungal phytohormones in pathogenic and mutualistic associations In:<br />

Carroll GC, Tudzynski P (eds) The mycota V. Springer, Berlin Heidelberg New York,<br />

pp 167–184<br />

Tudzynski B, Sharon A (2002) Biosynthesis, biological role and application of fungal hormones.<br />

In: Osiewacz HD (ed) The mycota X Industrial applications, Springer, Berlin<br />

Heidelberg New York, pp 183–211<br />

Usuki F, Narisawa K (2005) Formation of structures resembling ericoid mycorrhizas by<br />

the root endophytic fungus Heteroconium chaetospira within roots of Rhododendron<br />

obtusum var. kaempferi. Mycorrhiza 15:61–64<br />

Vandenkoornhuyse P, Baldauf SL, Leyval C, Straczek J, Young JPW (2002) Extensive fungal<br />

diversity in plant roots. Science 295:2051<br />

Varma A, Singh A, Sahay NS, Sharma J, Roy A, Kumari M, Raha D, Thakran S, Deka D,<br />

Bharti K, Hurek T, Blechert O, Rexer K-H, Kost G, Hahn A, Maier W, Walter M, Strack D,<br />

Kranner I (2000) Piriformospora indica: an axenically culturable mycorrhiza-like endosymbiotic<br />

fungus In: Hock B (ed) The mycota, vol IX. Fungal associations. Springer,<br />

Berlin Heidelberg New York, pp 125–150<br />

Verkley G (1999) A monograph of the genus Pezicula and its anamorphs. Stud Mycol<br />

44:5–171<br />

Villarreal-Ruiz L, Anderson IC, Alexander IJ (2004) Interaction between an isolate from<br />

the Hymenoscyphus ericae aggregate and roots of Pinus and Vaccinium New Phytol<br />

164:183–192<br />

Wilcox HE, Wang CJK (1987) Ectomycorrhizal and ectendomycorrhizal associations of<br />

Phialophora finlandia with Pinus resinosa, Picea rubens,andBetula alleghaniensis. Can<br />

J For Res 17:976–990<br />

Yates I, Bacon CW, Hinton DM (1997) Effects of endophytic infection by Fusarium moniliforme<br />

on corn growth and cellular morphology. Plant Dis 81:723–728<br />

Zhou D, Hyde K (2001) Host-specificity, host-exclusivity and host-recurrence in saprobic<br />

fungi. Mycol Res 105:1449–1457


2<br />

Spectrum and Population Dynamics<br />

of Bacterial Root Endophytes<br />

Johannes Hallmann, Gabriele Berg<br />

2.1<br />

Introduction<br />

Since the first reports regarding the existence of bacteria residing in plant<br />

roots (Trevet and Hollis 1948; Philipson and Blair 1957), numerous publicationshavedescribedthespectrumandpopulationdynamicsofindigenous<br />

endophytic root endophytes for various plant species (Bell et al. 1995;<br />

Gardner et al. 1982; Hallmann et al. 1997a, 1999; Hallmann 2003; Mahaffee<br />

and Kloepper 1997a, 1997b; McInroy and Kloepper 1995; Misaghi and<br />

Donndelinger 1990; Sturz et al. 1997). But what do we really know about<br />

the spectrum and population dynamics of those bacteria residing in the<br />

endorhiza? What are the potential risks associated with these bacteria?<br />

Answers to these questions will not only improve our understanding of<br />

plant/endophytic bacterial interactions, but are a prerequisite for any future<br />

commercialisation. This chapter reviews our current knowledge of<br />

(1) population density, bacterial spectrum and bacterial diversity of endophytic<br />

root bacteria, (2) factors influencing the population dynamics<br />

of indigenous and introduced bacterial endophytes, (3) bacterial interactions<br />

with biotic and abiotic factors, and (4) potential risks associated with<br />

endophytic bacteria.<br />

2.2<br />

Population Density<br />

Population densities of indigenous endophytic bacteria in roots are found<br />

to be about 10 5 cfu g −1 fresh root weight (Hallmann et al. 1997a). This is<br />

higher than in any other plant organ, as the population density usually<br />

decreases acropetally, with average densities of 10 4 cfu g −1 fresh weight in<br />

Johannes Hallmann: Federal Biological Research Centre for Agriculture and Forestry, Institute<br />

for Nematology and Vertebrate Research, Toppheideweg 88, 48161 Münster, Germany,<br />

E-mail: j.hallman@bba.de<br />

Gabriele Berg: Graz University of Biotechnology, Department of Environmental Technology,<br />

Petersgasse 12, 8010 Graz, Austria<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


16 J. Hallmann, G. Berg<br />

the stem and 10 3 cfu g −1 fresh weight in leaves. Generative organs such<br />

asflowers,fruitsandseedsarecolonisedbyevenlowernumbersand,in<br />

many cases, are below the detection limit. However, depending on plant<br />

species, methodology and other factors, reported population densities can<br />

vary significantly. Common densities reported for indigenous endophytic<br />

bacteria in roots range from 10 4 to 10 6 cfu g −1 for cotton and sweetcorn<br />

(McInroy and Kloepper 1994), 10 3 to 10 6 cfu g −1 for sugar beet (Jacobs et al.<br />

1985), 4.0×10 2 to 1.3×10 4 cfu g −1 for cotton (Hallmann et al. 1997a, Misaghi<br />

and Donndelinger 1990), 10 5 cfu g −1 for potato (Krechel et al. 2002) and<br />

10 5 cfu g −1 for pine seedlings (Shishido et al. 1995). Some authors have even<br />

reported population densities up to 10 10 cfu g −1 without negative effects on<br />

plant growth (Dimock et al. 1988; McInroy and Kloepper 1994). However,<br />

these high densities are considered exceptional; population densities above<br />

10 7 cfu g −1 are known frombacterial plant pathogens to cause pathogenicity<br />

(Tsiantos and Stevens 1986; Grimault and Prior 1994). Within the first<br />

3 weeks after seeding the population density generally increases to an<br />

optimum carrying capacity of about 10 5 cfu g −1 and then remains at this<br />

level for the rest of the growth period (Mahaffee and Kloepper 1997a, 1997b;<br />

McInroy and Kloepper 1995; Krechel et al. 2002). Even artificial inoculation<br />

of a highly concentrated bacterial suspension into the root tissue does not<br />

increase population density in the long run (Frommel et al. 1991). However,<br />

the above figures account only for culturable bacteria, which are only part<br />

of the total endophytic bacterial community. Their relative population size<br />

is still unknown as reliable data are lacking; but from other environments<br />

it is known that culturable bacteria represent between 0.001% and 15% of<br />

the total bacterial population.<br />

2.3<br />

Bacterial Spectrum<br />

Studying the bacterial spectrum requires bacterial identification. Before<br />

1990 this was laborious, using a combination of morphological and physiological<br />

characters often supported by ready-made tests such as API identification<br />

strips (Arrow Scientific, Lane Cove, Australia), or Biolog (Biolog,<br />

Hayward, CA) tests. With the fatty acid method described by Sasser (1990),<br />

bacterial identification and thus analysis of bacterial spectra became considerably<br />

more efficient, allowing time-course studies in which hundreds<br />

of bacteria could be identified (McInroy and Kloepper 1995; Mahaffee and<br />

Kloepper 1997a, 1997b). Nowadays, sequencing of 16S rDNA genes and<br />

alignment with public databases allows rapid and accurate species identification<br />

(Krechel et al. 2002; Reiter et al. 2002; Sessitsch et al. 2004). While<br />

this approach still relies on bacterial isolation (see Chap. 17 by Hallmann et


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 17<br />

al.), newly developed molecular methods use universal or specific primers<br />

to characterise either the endophytic community, certain bacterial groups<br />

or particular species, therefore enabling complete analysis of culturable as<br />

well as non-culturable bacteria (see Chap. 19 by van Overbeek et al.).<br />

As mentioned for bacterial density, the spectrum of indigenous endophytic<br />

bacteria is affected by similar biotic and abiotic factors. However, it<br />

also seems to be influenced by niche specialisation, differences in colonisation<br />

pathway (seed-borne, rhizosphere, above-ground plant organs), or<br />

some form of mutual exclusion operating between different bacterial populations<br />

(Hallmann 2001; Sturz et al. 1997). Table 2.1 gives an overview<br />

of endophytic bacteria commonly isolated from the roots of some cultivated<br />

plants. Although far from complete, it shows the broad spectrum of<br />

bacterial endophytes that occur in plant roots. Table 2.2 provides a more<br />

comprehensive list of bacterial species isolated from roots, comprising 219<br />

species representing 71 genera, with Bacillus and Pseudomonas being the<br />

most common genera.<br />

Table 2.1. Spectrum of endophytic bacteria most commonly isolated from roots of various<br />

cultivated plant species<br />

Plant Bacterial taxa Reference<br />

Alfalfa Pseudomonas, Erwinia-like bacteria Gagné et al. 1987<br />

Carrot Agrobacterium, Pseudomonas, Staphylococcus Surette et al. 2003<br />

Clover Agrobacterium, Bacillus, Methylobacterium,<br />

Pseudomonas, Rhizobium<br />

Sturz et al. 1997<br />

Cotton Bacillus, Burkholderia, Clavibacter, Erwinia, Chen et al. 1995;<br />

Phyllobacterium, Pseudomonas<br />

Hallmann et al. 1999;<br />

Misaghi and<br />

Donndelinger 1990<br />

Cucumber Agrobacterium, Bacillus, Burkholderia,<br />

Mahaffee and Kloepper<br />

Chryseobacterium, Clavibacter, Curtobacterium, 1997a, 1997b; McInroy<br />

Enterobacter, Micrococcus, Paenibacillus,<br />

Phyllobacterium, Pseudomonas, Serratia,<br />

Stenotrophomonas<br />

and Kloepper 1995<br />

Grapevine Enterobacter, Pseudomonas, Rahnella, Rhodococcus,<br />

Staphylococcus<br />

Bell et al. 1995<br />

Maize Agrobacterium, Arthrobacter, Bacillus, Burkholderia, Lalande et al. 1989;<br />

Corynebacterium, Curtobacterium, Enterobacter, McInroy and Kloepper<br />

Micrococcus, Phyllobacterium, Pseudomonas,<br />

Serratia<br />

1995<br />

Potato Agrobacterium, Arthrobacter, Bacillus,<br />

Krechel et al. 2002;<br />

Chryseobacterium, Enterobacter, Micrococcus,<br />

Pantoea, Pseudomonas, Stenotrophomonas,<br />

Streptomyces<br />

Sturz 1995


18 J. Hallmann, G. Berg<br />

Table 2.1. (continued)<br />

Canola Acidovorax, Agrobacterium, Aureobacterium,<br />

Bacillus, Cytophaga, Chryseobacterium,<br />

Flavobacterium, Micrococcus, Rathayibacter,<br />

Pseudomonas<br />

Red clover Agrobacterium, Bacillus, Methylobacterium, Pantoea,<br />

Pseudomonas, Rhizobium, Xanthomonas<br />

Germida et al. 1998;<br />

Graner et al. 2003;<br />

Misko and Germida<br />

2002<br />

Sturz et al. 1997<br />

Rice Serratia, Azoarcus Gyaneshwar et al. 2001;<br />

Hurek et al. 1994<br />

Rough<br />

lemon<br />

Bacillus, Corynebacterium, Enterobacter,<br />

Pseudomonas, Serratia<br />

Gardner et al. 1982<br />

Soybean Bacillus Bai et al. 2002<br />

Sugar beet Bacillus, Corynebacterium, Erwinia, Lactobacillus,<br />

Peudomonas, Xanthomonas<br />

Jacobs et al. 1985<br />

Sugar cane Acetobacter Cavalcante and<br />

Döbereiner 1988<br />

Tomato Bacillus, Burkholderia, Chryseobacterium, Kluyvera,<br />

Micrococcus, Pseudomonas, Serratia<br />

Munif 2001<br />

Wheat Bacillus, Flavobacterium, Microbispora, Micrococcus,<br />

Micromonospora, Nacardiodes, Rathayibacter,<br />

Streptomyces<br />

Diverse<br />

species<br />

Coombs and Franco<br />

2003;<br />

Germida et al. 1998<br />

Streptomyces Sardi et al. 1992<br />

Table 2.2. Endophytic bacteria isolated from roots of various plants<br />

Acetobacter diazotrophicus 13 , A. pasteurianus 9<br />

Acidovorax delafieldii 7,9,10 , A. facilis 10<br />

Acinetobacter baumannii 10 , A. calcoaceticus 10 , A. wolffi 5 , A. radioresistens 10<br />

Aeromonas salmonicida 9<br />

Agrobacterium radiobacter 7,8,9,10,12 , A. rhizogenes 12 , A. rubi 7,9 , A. tumefaciens 12,15<br />

Alcaligenes piechaudii 9,12 , A. xylosoxydans 9,12<br />

Arthrobacter atrocyaneus 10,11 , A. aurescens 11 , A. crystallopodietes 12 , A. ilicis 15 ,<br />

A. mysorens 10,12 , A. nicotianae 12 , A. pascens 10,11<br />

Aureobacterium barkeri12 , A. esteroaromaticum10,11 , A. liquefaciens11 , A. saperdae12 ,<br />

A. testaceum12 Azospirillum amazonense13 , A. brasiliense9,13 , A. lipoferum13 Bacillus alvei12 , B. amyloliquefaciens12 , B. azotoformans12,14 , B. brevis6,12,14,15 , B. cereus7,12 ,<br />

B. circulans10,14 , B. citinusporus11 , B. coagulans12,15 , B. fastidiosus2 , B. gordonae6 ,<br />

B. insolitus2,14 , B. laterosporus6,7,11,12 , B. lentus12 , B. licheniformis6 , B. longisporus6 ,<br />

B. macerans12 , B. megaterium6,7,10,11,12,14,15 , B. mycoides11 , B. pumilus6,7,11,12 ,<br />

B. sphaericus7,12 , B. subtilis7,12,14 , B. throphaeus6 , B. thuringiensis12


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 19<br />

Table 2.2. (continued)<br />

Bordetella avium 14 , B. bronchiseptica 8<br />

Brevibacterium acetylicum 11<br />

Brevundimonas diminuta 9 , B. vesicularis 7,9<br />

Burkholderia cepacia 7,8,9,12 , B. gladioli 8,10,12 , B. pickettii 7,8,9,12 , B. solanacearum 12<br />

Cellulomonas cellulans 12 , C. flavigena 10 , C. turbata 12,15<br />

Chryseobacterium balustinum 9,11 , C. indologenes 7,11 , C. meningosepticum 7,11<br />

Citrobacter freundii 5 , C. koseri 12<br />

Clavibacter michiganensis 2,7,11,12,15<br />

Comamonas acidovorans 7,8,9,10,11 , C. terrigena 2 , C. testosteroni 9,12,14<br />

Curtobacterium allidum15 , C. citreum14 , C. flaccumfaciens2,10,12,14,15 , C. luteum14,15 ,<br />

C. pusillum2,12,15 Cytophaga aquatilis10 , C. johnsonae9,10,11 Enterobacter aerogenes 5 , E. asburiae 8,10,12 , E. cancerogenus 12 , E. cloacae 2,5,8,12 ,<br />

E. intermedius 11 , E. taylorae 7,8,12 , E. sakazakii 5<br />

Erwinia carotovora 9,12 , E. chrysanthemi 9<br />

Escherichia coli 12,14 , E. vulneris 12<br />

Flavimonas oryzihabitans 12<br />

Flavobacterium aquatile 11 , F. indologenes 12 , F. meningosepticum 12 , F. resinovorum 9,10,11<br />

Gluconobacter oxidans 9<br />

Gordonia polyisoprenivorans 3<br />

Herbaspirillum spp. 13<br />

Hydrogenophaga flava 12 , H. pseudoflava 10,12<br />

Kingella kingae 15<br />

Klebsiella ozaenae 2 , K. planticola 12 , K. pneumoniae 2,12 , K. terrigena 12<br />

Kluyvera ascorbata 12 , K. cryocrescens 8,9,10,12<br />

Kocuria kristinae 11<br />

Lactobacillus kefir 10<br />

Leuconostoc pseudomesenteroides 10<br />

Methylobacterium extorquens14 , M. fuijsawaense12 , M. mesophilicum12 , M. radiotolerans12 ,<br />

M. rhodinum14 Microbacterium imperiale12 , M. laevaniformans12 Micrococcus agilis 12 , M. halobius 9,10,11 , M. kristinae 12 , M. luteus 8,9,10,11,12 , M. lylae 10,12 ,<br />

M. roseus 12 , M. varians 10,12,15<br />

Micromonospora endolithica 3 , M. peucetica 3<br />

Moraxella bovis 2 , M. phenylpyruvica 10<br />

Mycobacterium aichiense3 , M. bohemicum3 , M. cookii3 , M. flavescens3 , M. heidelbergense3 ,<br />

M. palustre3 Neisseria flavescens, N. mucosa9


20 J. Hallmann, G. Berg<br />

Table 2.2. (continued)<br />

Nocardia pseudobrasiliensis 3<br />

Nocardiodes albus 4<br />

Ochrabactrum anthropi 12<br />

Paenibacillus alvei 10 , P. pabuli 12 , P. polymyxa 12<br />

Pantoea agglomerans 2,5,11,12,14,15 , P. ananas 12 , P. stewartii 9<br />

Paracoccus denitrificans 7<br />

Pedicoccus pentosaceus 9<br />

Photobacterium angustum 9<br />

Phyllobacterium myrsinacearum 7,8,9,12,14 , P. rubiacearum 7,8,9,10,12<br />

Plesiomonas shigelloides 10<br />

Providencia spp. 5<br />

Pseudomonas aeruginosa 5,9 , P. cepacia 5 , P. chlororaphis 6,7,10,11,12 , P. cichorii 2,6,12,15 ,<br />

P. coronofaciens 12 , P. corrugata 2,6,10,14,15 , P. flectens 10 , P. fluorescens 5,6,7,9,11,12 , P. fragi 14 ,<br />

P. fulva 14 , P. marginalis 2,6,9,12 , P. mendocina 6,9,12 , P. pseudoalcaligenes 6 ,<br />

P. putida 2,5,6,7,9,11,12 , P. rubrisubalbicans 9,12 , P. saccharophila 8,10,12 , P. savastanoi 6,10,12 ,<br />

P. stutzeri 9 , P. syringae 2,6,9,11,12 , P. tolaasii 14,15 , P. vesicularis 12,14 , P. viridiflava 6,11<br />

Rahnella aquatilis 2<br />

Rhizobium etli 10 , R. japonicum 12 , R. leguminosarum 14 , R. loti 14 , R. meliloti 15<br />

Rhodococcus coprophilus 3 , R. luteus 2<br />

Salmonella cholerasuis 7<br />

Serratia liquefaciens 5,12 , S. marcescens 12 , S. plymuthica 12<br />

Shigella spp. 5<br />

Shingobacterium heparinum 11<br />

Sphingomonas capsulata 9 , S. paucimobilis 9,12,15 , S. thalpophilum 15<br />

Staphylococcus capitis 12 , S. epidermidis 12 , S. haemolyticus 11<br />

Stenotrophomonas maltophilia 9,11,12<br />

Streptomces argenteolus 4 , S. bikiniensis 4 , S. caviscabies 4 , S. cyaneus 11 , S. galilaeus 4 ,<br />

S. halstedii 11 , S. maritimus 4 , S. pseudovenezuelae 4 , S. scabies 4 , S. setonii 4 , S. tendae 4 ,<br />

S. thermolineatus 3 , S. violaceusniger 11<br />

Variovorax campestris, V. paradoxus 7,9,12<br />

Vibrio cholerae 9<br />

Xanthobacter agilis 9,10<br />

Xanthomonas agilis 7 , X. campestris 2,8,9,12,14,15 , X. oryzae 15<br />

Yersinia frederiksenii 12 , Y. pseudotuberculosis 10<br />

a <strong>References</strong>:<br />

1 Adhikari et al. 2001; 2 Bell et al. 1995; 3 Conn and Franco 2004; 4 Coombs and Franco 2003;<br />

5 Gardner et al. 1982; 6 Germida and Siciliano 2001; 7,8,9 Hallmann et al. 1997b, 1998, 1999;<br />

10 Hallmann 2003; 11 Krechel et al. 2002; 12 McInroy and Kloepper 1995; 13 Reis et al. 2000;<br />

14,15 Sturz et al. 1997, 1999


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 21<br />

2.4<br />

Bacterial Diversity<br />

Diversity indices allow the compression of bacterial spectrum and bacterial<br />

density into a single number to facilitate comparisons between plant<br />

species, habitats, etc., as well as elucidation of changes in community relationships<br />

(Mahaffee and Kloepper 1997a, 1997b). Of the many diversity<br />

indices used in ecology, only few can be applied to endophytic bacteria.<br />

Three of the more commonly used indices are genera richness, Hill’s modified<br />

Shannon’s index N1, and Hill’s modified Simpson’s index N2, with N2<br />

more than N1 representing very abundant species (Ludwig and Reynolds<br />

1988). Using these indices for the endorhiza of field-grown cucumber, Mahaffee<br />

and Kloepper (1997a, 1997b) observed that all three indices tended to<br />

increase over the growing season, reaching their highest values at the final<br />

sampling date 70 days after planting. However, density varied between two<br />

consecutive years, indicating the importance of climatic conditions for the<br />

community structure of bacterial root endophytes. For potatoes, Krechel et<br />

al. (2002) observed the highest bacterial diversity at flowering. During that<br />

period the plant undergoes massive physiological changes that probably<br />

increase nutrient availability and thus bacterial diversity. Other biotic and<br />

abiotic factors that influence bacterial diversity are discussed later in this<br />

chapter. There is little question, though, that a better understanding of the<br />

population dynamics of bacterial root endophytes will enhance our ability<br />

to take advantage of their beneficial potential to enhance plant growth and<br />

health.<br />

2.5<br />

Factors Influencing Colonisation<br />

The enormous spectrum and high diversity of endophytic bacteria found<br />

in different plant species begs the question: what biotic and abiotic factors<br />

influence bacterial colonisation?<br />

2.5.1<br />

Methodology<br />

Methodology is especially important in describing the spectrum of culturable<br />

bacteria, as different methods will give different results. Key factors<br />

affecting the bacterial spectrum recovered are (1) length of surface sterilisation,<br />

(2) concentration of the sterilising agent and (3) the method itself.<br />

Comparison of the trituration method with the pressure bomb technique


22 J. Hallmann, G. Berg<br />

(see Chap. 17 by Hallmann et al.) revealed significantly larger populations<br />

of endophytic bacteria in cotton roots using the first method (Hallmann et<br />

al. 1997b). However, total number of bacterial genera and species recovered<br />

was greater using the pressure bomb technique, mainly because a higher<br />

number of less commonly occurring species was recovered. These results<br />

suggest that the two techniques sample two different microhabitats, i.e.<br />

the pressure bomb technique more effectively recovering mainly vascular<br />

colonists while the trituration method recovers both vascular colonisers<br />

and bacteria residing in the root cortex. An even higher bacterial spectrum<br />

can be identified using molecular methods that detect culturable as well as<br />

non-culturable bacteria (see Chap. 17 by Hallmann et al.).<br />

2.5.2<br />

Geography<br />

Geographical regions are believed to differ in their bacterial spectra. Munif<br />

(2001) compared the bacterial spectrum of tomato roots grown under temperate<br />

conditions in Bonn, Germany, with that of tomato roots grown under<br />

tropical conditions in Bogor, Indonesia. In Germany, 38 species comprising<br />

21 genera were isolated, whereas in Indonesia, 50 species comprising<br />

32 genera were isolated. Twenty-four bacterial species were exclusively isolated<br />

from tomato roots in Germany and 38 species exclusively from tomato<br />

roots in Indonesia. However, 14 species were isolated from both regions.<br />

Interestingly, the most abundant species in both geographical regions were<br />

Pseudomonas putida and Bacillus megaterium. These two bacterial species<br />

are also commonly reported in other regions of the world (Tables 2.1, 2.2).<br />

Howisitpossiblethatabacterialspeciescolonisessuchdifferentregions<br />

and still dominates the endophytic spectrum? And are populations from<br />

different regions distinguishable? Using molecular fingerprint techniques,<br />

fluorescent Pseudomonas strains collected worldwide could be regionally<br />

grouped suggesting that adaptation to their specific environment has occurred<br />

(Cho and Tiedje 2000). But adaptations also seem to occur within<br />

a given environment. For example, Pseudomonas fluorescens strains isolated<br />

from the endorhiza of potato are distinguishable from P. fluorescens<br />

strains isolated from the rhizosphere of the same plant (Berg et al. 2005).<br />

2.5.3<br />

Plant Species<br />

The effect of plant species on the bacterial spectrum being recovered has<br />

only been marginally studied. Plant specificity of the bacterial community<br />

was shown for the rhizosphere by Smalla et al. (2001) and Berg et al. (2002).


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 23<br />

As endophytic bacteria represent a subset of the rhizosphere colonisers,<br />

one would expect that different plant species growing side by side would be<br />

colonised by a different spectrum of endophytic bacteria. This hypothesis<br />

was confirmed by McInroy and Kloepper (1995), who showed that the bacterial<br />

spectrum of sweet corn and cotton grown next to each other differed.<br />

Although the total number of genera was similar for both plant species,<br />

some genera, such as Alcaligenes, Aureobacterium, Cellulomonas, Comamonas,<br />

Erwinia, Ochrobactrum, andYersinia, were recovered only from<br />

cotton roots, while other genera, such as Arthrobacter, Citrobacter, Flavimonas,<br />

Microbacterium and Stenotrophomonas, were recovered exclusively<br />

from sweetcorn roots. Plant-species-specific factors such as root architecture,<br />

surface structure, and composition of the root exudates as well as<br />

non-plant factors such as mycorrhization or wounding probably influence<br />

the bacterial spectrum prior to colonisation. Following root colonisation,<br />

size of the intercellular space, nutrient composition within the apoplastic<br />

fluid, and the plant’s response to endophytic colonisation are probably<br />

the main factors determining bacterial selectivity and thus the bacterial<br />

spectrum found inside the roots.<br />

2.5.4<br />

Plant Genotype<br />

Does the plant genotype affect the spectrum of endophytic bacteria? What<br />

about cultivars resistant to bacterial plant pathogens? For the latter, no<br />

differences were seen in the population densities of endophytic bacteria of<br />

two grapevine cultivars resistant and susceptible to Agrobacterium tumefaciens,<br />

the causal agent of crown gall (Bell et al. 1995). However, resistance<br />

varied in some colonised cultivars (Conn et al. 1997), suggesting that not<br />

only the host genotype but also the associated bacterial endophytes may<br />

contribute to plant resistance, as suggested by Bird et al. (1980). Differences<br />

in the endophytic spectrum were also reported to occur in potato<br />

tubers of four potato cultivars (Sturz et al. 1999). Although the two bacterial<br />

species Curtobacterium flaccumfaciens and Pseudomonas cichorii occurred<br />

in all four cultivars, cultivar-specific preferences were observed.<br />

For example, C. flaccumfaciens dominated the endophytic spectrum of<br />

potato ‘Kennebec’ but was rarely found in ‘Butte’, whereas P. cichorii was<br />

dominant in ‘Butte’ and less frequently isolated from the other three cultivars.<br />

In addition, some bacterial species were exclusively isolated from<br />

one cultivar, but not from others. Similar to plant species, plant genotypes<br />

also vary in their biochemical composition, which may thus affect bacterial<br />

spectra. Similarly, Graner et al. (2003) found noticeable differences in<br />

endophytic bacterial populations of oil-seed rape cultivars that differ in


24 J. Hallmann, G. Berg<br />

their susceptibility to the fungal wilt pathogen Verticillium longisporum.<br />

Genotype-specific differences were also shown for wheat cultivars: modern<br />

cultivars supported a more diverse endophytic community than ancient<br />

land races (Germida and Siciliano 2001). Furthermore, endophytic bacteria<br />

isolated from different canola cultivars had different carbon utilisation<br />

profiles (Misko and Germida 2002).<br />

2.6<br />

Interactions<br />

Ecological theory states that when a disruptive force such as pathogen<br />

infestation affects a community, the diversity of that community initially<br />

increases and its membership becomes highly variable before reaching<br />

a new equilibrium (Mahaffee and Kloepper 1997a; Petratis et al. 1989). But<br />

does this also apply to endophytic bacteria of roots? Plant roots are continuously<br />

challenged by soil-borne pathogens, mutualistic symbionts and<br />

various soil conditions. As a result, plant defence mechanisms might be induced.<br />

How do these parameters affect the endophytic bacterial spectrum?<br />

2.6.1<br />

Plant Pathogens<br />

While the beneficial effects of endophytic bacteria to control plant pathogens<br />

are well documented [see Chaps. 3 (Kloepper and Ryu) and 4 (Berg<br />

and Hallmann)], very little is known about how plant pathogens affect<br />

the spectrum and diversity of bacterial root endophytes. Regarding fungal<br />

pathogens, Mahaffee et al. (1997a, 1997b) reported that root infection<br />

by Rhizoctonia solani promoted colonisation of the two introduced bacterial<br />

endophytes Enterobacter asburiae JM22 and Pseudomonas fluorescens<br />

89B-27. Interactions between plant pathogenic bacteria and endophytic<br />

bacteria have so far been described only for above ground plant organs.<br />

Inoculation of potatoes with Erwinia carotovora subsp. atroseptica caused<br />

an increase in endophytic bacterial diversity of infected plants compared<br />

with healthy plants (Reiter et al. 2002); similarly, infestation of citrus by<br />

Xylella fastidiosa was positively correlated with the occurrence of Methylobacterium<br />

spp. (Araújo et al. 2002). A third group of plant pathogens,<br />

plant parasitic nematodes, significantly increased the total number of bacterial<br />

root endophytes (Hallmann et al. 1998; Hallmann 2003), and also<br />

affected the bacterial spectrum (Hallmann 2003). In cotton roots infested<br />

with the root-knot nematode Meloidogyne incognita, 11 species were exclusively<br />

isolated from roots of infested plants, while 15 species occurred<br />

only in non-infested plants.


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 25<br />

2.6.2<br />

Plant Symbionts<br />

Interactions between endophytic bacteria and mutualistic plant symbionts<br />

have been best studied for nodule bacteria. Sturz et al. (1997) reported that<br />

bacterial density and diversity was lower in root nodules of red clover than<br />

in tap roots, but root nodules yielded a higher number of species exclusively<br />

colonising this specific niche (Sturz et al. 1997). Coinoculation of<br />

Rhizobium leguminosarum BV trifolii with endophytic bacteria promoted<br />

root nodulation. In other cases, endophytic bacteria reduced or inhibited<br />

root colonisation by nitrogen-fixing bacteria (Bacilio-Jiménez et al. 2001),<br />

indicating that those interactions seem to be strain-specific. Little is yet<br />

known about the interaction between endophytic bacteria and other symbionts<br />

such as fungal endophytes and mycorrhizal fungi. Mycorrhizal fungi<br />

are at least known to influence rhizosphere bacteria (Azacón-Aguilar and<br />

Barea 1992) and similar effects might apply to endophytic bacteria.<br />

2.6.3<br />

Plant Defence Mechanisms<br />

A question frequently asked is: why are endophytic bacteria not inhibited<br />

by a plant defence response? And if there is a plant defence response,<br />

how does this affect the bacterial spectrum? To answer these questions,<br />

potato plant resistance was experimentally induced by Rhizobium etli G12<br />

(Hallmann 2003). R. etli G12 was applied to one-half of a split potato root<br />

system and the endophytic bacterial spectrum was analysed for the other<br />

(“induced”) half of the split root system and compared with that of nontreated<br />

plants. Total bacterial density was significantly higher in treated<br />

(1.6×10 4 cfu g −1 ) than in non-treated (3.2×10 3 cfu g −1 ) roots. Seventeen<br />

bacterial species were isolated from treated roots compared with 14 species<br />

from non-treated roots. The results indicated that, under the conditions<br />

of bacteria-mediated induced plant defence responses, the density and<br />

spectrumofbacterialrootendophytesisincreased.<br />

2.6.4<br />

Agricultural Practices<br />

Besides the plant itself and plant-associated microorganisms, agricultural<br />

practices can also influence the spectrum and population dynamics of bacterial<br />

root endophytes. Seghers et al. (2004) showed that different fertiliser<br />

treatments influenced the endophytic community of maize roots, whereas


26 J. Hallmann, G. Berg<br />

the tested herbicide treatments did not. High N-fertilisation inhibited the<br />

colonisation of sugarcane by Acetobacter diazotrophicus (Fuentes-Ramírez<br />

et al. 1999), while application of nitrogen-containing chitin as an organic<br />

amendment supported endophytic species in cotton roots that otherwise<br />

did not occur (Hallmann et al. 1999). For the latter, it was shown that<br />

the endophytic spectrum was completely different from that of the rhizosphere,<br />

indicating that the bacterial composition of the rhizosphere is not<br />

the only factor determining the endophytic spectrum. Differences in plant<br />

biochemistry due to the organic amendment, such as enhanced chitinase<br />

and peroxidase concentrations (Hallmann 2003), might have changed the<br />

plants’ preference for certain bacterial endophytes.<br />

2.7<br />

Potential Human Pathogens Among Root Endophytes<br />

By definition, endophytic bacteria reside within the plant without causing<br />

visible disease symptoms (see Chap. 1 by Schulz and Boyle). However, the<br />

distinction between harmless and harmful bacteria is not always clear.<br />

Potentially harmful bacteria might colonise the plant latently or reside as<br />

dormant stages, becoming harmful only at later growth stages or when<br />

a critical density, known as “quorum sensing”, is reached (Eberl 1999).<br />

Since the beneficial attributes of endophytic bacteria are covered elsewhere<br />

in this book [see Chaps. 3 (Kloepper and Ryu), 4 (Berg and Hallmann),<br />

and 6 (Anand et al.)], we will focus on potentially human pathogenic<br />

bacteria. In general, mechanisms of pathogenicity and antagonism are<br />

very similar, and sometimes only the expression of one metabolite or total<br />

bacterial numbers dictates whether the bacteria are harmful or harmless<br />

(Suckstorff and Berg 2003; Berg et al. 2006). For example, type III secretion<br />

systems, known for their role in bacterial pathogenicity, are present in<br />

many plant-associated Pseudomonas strains and may confer induced resistance,<br />

plant growth promotion or biological control (Preston et al. 2001).<br />

A few bacterial species isolated from inside plant roots are also known to be<br />

pathogenic to humans These include strains of Bacillus cereus, Burkholderia<br />

cepacia, Serratia marcescens and Stenotrophomonas maltophilia,whichnot<br />

only have excellent antagonistic properties against plant pathogens, but are<br />

also known to cause human diseases, especially of debilitated or immunosuppressed<br />

individuals (LiPuma 2003; Minkwitz and Berg 2001; Vandamme<br />

and Mahenthiralingam 2003; Wolf et al. 2002). Staphylococcus species, also<br />

known to be human pathogens, have been isolated from the potato endorhiza<br />

(Krechel et al. 2002; Reiter et al. 2002). In some cases, humanpathogenic<br />

isolates differ from the harmless isolates occurring in plants by<br />

the expression of certain toxins, in other cases no such distinction has yet


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 27<br />

been found. Reiter et al. (2002) showed, based on 16S rDNA sequences, high<br />

homology of endophytic bacterial isolates from potato leaves with those of<br />

human pathogens such as Enterobacter amnigenus, Enterobacter cloacae,<br />

Stenotrophomonas maltophilia, Staphylococcus xylosus and Ochrobactrum<br />

anthropi. E. cloacae, S. maltophilia,andO. anthropi have also been reported<br />

to occur in plant roots (McInroy and Kloepper 1995).<br />

Nevertheless, the presence of human-pathogenic bacteria in plant roots<br />

can be a health concern and therefore should be carefully monitored. As<br />

a result of several Salmonella outbreaks attributed to alfalfa sprout contamination<br />

in Finland and the United States (van Beneden et al. 1999), alfalfa<br />

seeds were found to be contaminated and surface sterilisation did not<br />

eliminate the enteric pathogen. Gandhi et al. (2001) detected a Salmonella<br />

Stanley strain inside alfalfa sprouts to a depth of 18 µm without finding cells<br />

on the surface, thus confirming the endophytic properties of this pathogen.<br />

Strains isolated from patients infected with Salmonella enterica during an<br />

alfalfa-sprout-associated outbreak and labelled with the green fluorescent<br />

protein (GFP) successfully colonised alfalfa seedlings (Dong et al. 2003). An<br />

inoculation with very few cells was sufficient to colonise the plant interior;<br />

however, strains of S. enterica differed greatly in their ability to invade the<br />

plant interior and to colonise alfalfa roots. This demonstrates the need for<br />

further investigations to ensure food safety in the future.<br />

2.8<br />

Conclusions<br />

In conclusion, of the plants thus far studied, the spectrum and diversity of<br />

endophytic bacteria in the roots varies greatly. What about the endophytic<br />

bacterial spectrum of plants growing under extreme climatic conditions,<br />

such as halophytes and xerophytes? Survival mechanisms developed by<br />

those bacteria may have some interesting industrial or pharmaceutical applications.<br />

Newly developed cultivation-independent methods have made<br />

clear that there is much more diversity among endophytic bacteria than at<br />

first expected. The major factors influencing bacterial diversity and colonisation<br />

have been discussed and their potential to manage endophytic communities<br />

towards increased benefits for plants and human health have been<br />

outlined. However, the potential risks of endophytic bacteria, especially of<br />

those strains known also to be potential human pathogens, need further<br />

exploration.


28 J. Hallmann, G. Berg<br />

<strong>References</strong><br />

Adhikari TB, Joseph CM, Yang G, Phillips DA, Nelson LM (2001) Evaluation of bacteria<br />

isolated from rice for plant growth promotion and biological control of seedling disease<br />

of rice. Can J Microbiol 47:916–924<br />

Araújo WL, Marcon J, Maccheroni W Jr, Van Elsas JD, Van Vuurde JW, Azevedo JL (2002) Diversity<br />

of endophytic bacterial populations and their interaction with Xylella fastidiosa<br />

in citrus plants. Appl Environ Microbiol 68:4906–4914<br />

Azacón-Aguilar C, Barea JM (1992) Interactions between mycorrhizal fungi and other<br />

rhizosphere microorganisms. In: Allen MF (ed) Mycorrhizal functioning: an integrative<br />

plant-fungal process. Chapmann & Hall, New York, pp 163–198<br />

Bacilio-Jiménez M, Aguilar-Flores S, del Valle MV, Pérez A, Zapeda A, Zenteno E (2001)<br />

Endophytic bacteria in rice seeds inhibit early colonization of roots by Azospirillum<br />

brasilense. Soil Biol Biochem 33:167–172<br />

Bai Y, D’aoust F, Smith DL, Driscoll BT (2002) Isolation of plant-growth-promoting Bacillus<br />

strains from soybean root nodules. Can J Microbiol 48:230–238<br />

Bell CR, Dickie GA, Harvey WLG, Chan JWYF (1995) Endophytic bacteria in grapevine.<br />

Can J Microbiol 41:46–53<br />

Berg G, Roskot N, Steidle A, Eberl L, Zock A, Smalla K (2002) Plant-dependent genotypic and<br />

phenotypic diversity of antagonistic rhizobacteria isolated from different Verticillium<br />

host plants. Appl Environ Microbiol 68:3328–3338<br />

BergG,Krechel A,Ditz M,Sikora RA,Ulrich A,Hallmann J(2005) Endophyticand ectophytic<br />

potato-associated bacterial communities differ in structure and antagonistic function<br />

against plant pathogenic fungi. FEMS Microbiol Ecol 51:215–229<br />

Berg G, Eberl L, Hartmann A (2006) The rhizosphere as a reservoir for opportunistic human<br />

pathogenic bacteria. Environ. Microbiol. 71:4203–4213<br />

Bird LS, Leverman C, Thaxaton R, Percy RG (1980) Evidence that microorganisms in and on<br />

tissues have a role in a mechanism of multi-adversity resistance in cotton. In: Brown JM<br />

(ed) Proceedings of Beltville Cotton Production Research Conferences. Nationals Cotton<br />

Council, Memphis, TN, pp 283–285<br />

Cavalcante VA, Döbereiner J (1988) A new acid-tolerant nitrogen fixing bacterium associated<br />

with sugarcane. Plant Soil 108:23–31<br />

Chen C, Bauske EM, Musson G, Rodríguez-Kábana R, Kloepper JW (1995) Biological control<br />

of Fusarium wilt on cotton by use of endophytic bacteria. Biol Control 5:83–91<br />

Cho JC, Tiedje JM (2000) Biogeography and degree of endemicity of fluorescent Pseudomonas<br />

strains in soil. Appl Environ Microbiol 66:5448–5456<br />

Conn, VM, Franco CMM (2004) Analysis of the endophytic actinobacterial population<br />

in the roots of wheat (Triticum aestivum L.) by terminal restriction fragment length<br />

polymorphism and sequencing on 16S rRNA clones. Appl Envrion Microbiol 70:1787–<br />

1794<br />

Conn KL, Nowak J, Lazorovita G (1997) A gnotobiotic bioassay for studying interactions between<br />

potatoes and plant growth-promoting rhizobacteria. Can J Microbiol 43:801–808<br />

Coombs JT, Franco CMM (2003) Isolation and identification of actinobacteria from surfacesterilized<br />

wheat roots. Appl Environ Microbiol 69:5603–5608<br />

Dimock MB, Beach RM, Carlson PS (1988) Endophytic bacteria for delivery of crop protection<br />

agents. In: Roberts DW, Granados RR (eds) Biotechnology, biological pesticides<br />

and novel plant-pest resistance for pest management. Boyce Thompson Institute for<br />

Plant Research, Ithaca, NY, pp 88–92<br />

Dong Y, Iniguez AL, Ahmer BM, Triplett EW (2003) Kinetics and strain specificity of<br />

rhizosphere and endophytic colonization by enteric bacteria on seedlings of Medicago<br />

sativa and Medicago truncatula. Appl Environ Microbiol 69:1783–1790


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 29<br />

Eberl L (1999) N-acyl homoserinelactone-mediated gene regulation in Gram-negative bacteria.<br />

Syst Appl Microbiol 22:493–506<br />

Frommel MI, Nowak J, Lazarovits G (1991) Growth enhancement and developmental modificationsofinvitrogrownpotato(Solanum<br />

tuberosum ssp. tuberosum) asaffectedby<br />

anonfluorescentPseudomonas sp. Plant Physiol 96:928–936<br />

Fuentes-Ramírez LE, Caballero-Melado J, Sepúlveda J, Martínez-Romero E (1999) Colonization<br />

of sugarcane by Acetobacter diazotrophicus is inhibited by high N-fertilization.<br />

FEMS Microbiol Ecol 29:117–128<br />

Gagné S, Richard C, Rousseau H, Antoun H (1987) Xylem-residing bacteria in alfalfa roots.<br />

Can J Microbiol 33:996–1000<br />

Gandhi MS, Golding S, Yaron S, Matthews KR (2001) Use of green fluorescent protein<br />

expressing Salmonella Stanley to investigate survival, spatial location, and control on<br />

alfalfa sprouts. J Food Prot 64:1891–1898<br />

Gardner JM, Feldman AW, Zablotowicz RM (1982) Identity and behavior of xylem-residing<br />

bacteria in rough lemon roots of Florida citrus trees. Appl Environ Microbiol 43:1335–<br />

1342<br />

Germida JJ, Siciliano SD (2001) Taxonomic diversity of bacteria associated with the roots<br />

of modern, recent and ancient wheat cultivars. Biol Fertil Soils 33:410–415<br />

Germida JJ, Siciliano SD, Freitas JR de, Seib AM (1998) Diversity of root-associated bacteria<br />

associated with field-grown canola (Brassica napus L.) and wheat (Tricicum aestivum<br />

L.). FEMS Microbiol Ecol 26:43–50<br />

Graner G, Persson P, Meijer J, Alstrom S (2003) A study on microbial diversity in different<br />

cultivars of Brassica napus in relation to its wilt pathogen, Verticillium longisporum.<br />

FEMS Microbiol Lett 29:269–276<br />

Grimault V, Prior P (1994) Invasiveness of Pseudomonas solanacearum in tomato, eggplant,<br />

and pepper: a comparative study. Eur J Plant Pathol 100:259–267<br />

Gyaneshwar P, James EK, Mathan N, Reddy PM, Reinhold-Hurek B, Ladha JK (2001) Endophytic<br />

colonization of rice by a diazotrophic strain of Serratia marcescens. JBacteriol<br />

183:2634–2645<br />

Hallmann J (2001) Plant interactions with endophytic bacteria. In: Jeger MJ, Spence NJ (eds)<br />

Biotic interactions in plant-pathogen associations. CABI, Wallingford, UK, pp 87–119<br />

Hallmann J (2003) Biologische Bekämpfung pflanzenparasitärer Nematoden mit antagonistischen<br />

Bakterien, vol 392. Mitt Biol Bundesanst Land Forstwirtsch, Berlin,<br />

Hallmann J, Quadt-Hallmann A, Mahaffee WF, Kloepper JW (1997a) Bacterial endophytes<br />

in agricultural crops. Can J Microbiol 43:895–914<br />

Hallmann J, Kloepper JW, Rodríguez-Kábana R (1997b) Application of the Scholander<br />

pressure bomb to studies on endophytic bacteria of plants. Can J Microbiol 43:411–416<br />

Hallmann J, Quadt-Hallmann A, Rodríguez-Kábana R, Kloepper JW (1998) Interactions<br />

between Meloidogyne incognita and endophytic bacteria in cotton and cucumber. Soil<br />

Biol Biochem 30:925–937<br />

Hallmann J, Rodríguez-Kábana R, Kloepper JW (1999) Chitin-mediated changes in bacterial<br />

communities of the soil, rhizosphere and within roots of cotton in relation to nematode<br />

control. Soil Biol Biochem 31:551–560<br />

Hurek T, Reinhold-Hurek B, van Montagu M, Kellenberger E (1994) Root colonization and<br />

systemic spreading of Azoarcus sp. Strain BH72 in grasses. J Bacteriol 176:1913–1923<br />

Jacobs MJ, Bugbee WM, Gabrielson DA (1985) Enumeration, location, and characterization<br />

of endophytic bacteria within sugar beet roots. Can J Bot 63:1362–1365<br />

Krechel A, Faupel A, Hallmann J, Ulrich A, Berg G (2002) Potato-associated bacteria and their<br />

antagonistic potential towards plant-pathogenic fungi and the plant-parasitic nematode<br />

Meloidogyne incognita (Kofoid & White) Chitwood. Can J Microbiol 48:772–786


30 J. Hallmann, G. Berg<br />

Lalande R, Bissonnette N, Coutlée D, Antoun H (1989) Identification of rhizobacteria from<br />

maize and determination of their plant-growth promoting potential. Plant Soil 115:7–11<br />

LiPuma JJ (2003) Burkholderia cepacia complex as human pathogens. J Nematol 35:213–217<br />

Ludwig JA, Reynolds JF (1988) Statistical ecology. Wiley, New York<br />

Mahaffee WF, Kloepper JW (1997a) Temporal changes in the bacterial communities of soil,<br />

rhizosphere, and endorhiza associated with field-grown cucumber (Cucumis sativus L.).<br />

Microb Ecol 34:210–223<br />

Mahaffee WF, Kloepper JW (1997b) Bacterial communities of the rhizosphere and endorhiza<br />

associated with field-grown cucumber plants inoculated with a plant growth-promoting<br />

rhizobacterium or its genetically modified derivative. Can J Microbiol 43:34–35<br />

McInroy JA, Kloepper JW (1994) Studies on indigenous endophytic bacteria of sweetcorn<br />

and cotton. In: O’Gara F, Dowling DN, Boesten B (eds) Molecular ecology of rhizosphere<br />

microorganisms. VCH, Weinheim, pp 19–28<br />

McInroy JA, Kloepper JW (1995) Survey of indigenous bacterial endophytes from cotton<br />

and sweet corn. Plant Soil 173:337–342<br />

Minkwitz A, Berg G (2001) Comparison of antifungal activities and 16S ribosomal DNA<br />

sequences of clinical and environmental isolates of Stenotrophomonas maltophilia.JClin<br />

Microbiol 39:139–145<br />

Misaghi IJ, Donndelinger CR (1990) Endophytic bacteria in symptom-free cotton plants.<br />

Phytopathology 80:808–811<br />

Misko AL, Germida JJ (2002) Taxonomic and functional diversity of pseudomonads isolated<br />

from the roots of field-grown canola. FEMS Microbiol Ecol 42:399–407<br />

Munif A (2001) Studies on the importance of endophytic bacteria for the biological control<br />

of the root-knot nematode Meloidogyne incognita on tomato. PhD thesis, University of<br />

Bonn, Germany<br />

Petratis PS, Latham RE, Niesenbaum RA (1989) The maintenance of species diversity by<br />

disturbance. Q Rev Biol 64:393–418<br />

Philipson MN, Blair ID (1957) Bacteria in clover root tissue. Can J Microbiol 3:135–139<br />

Preston GM, Bertrand N, Rainey PB (2001) Type III secretion in plant growth-promoting<br />

Pseudomonas fluorescens SBW25. Mol Microbiol 41:999–1014<br />

Reis Junior dos FB, Silva da LG, Reis VM, Dobereiner J (2000) Occurrence of diazotrophic<br />

bacteria in different sugar cane genotypes. Pesqui Agropecu Bras 35:985–994<br />

Reiter B, Pfeifer U, Schwab H, Sessitsch A (2002) Response of endophytic bacterial communities<br />

in potato plants to infection with Erwinia carotovora subsp. atroseptica.Appl<br />

Environ Microbiol 68:2261–2268<br />

Sardi P, Sarachhi M, Quaroni S, Petrolini B, Borgonovi GE, Merli S (1992) Isolation of<br />

endophytic Streptomyces strains from surface-sterilized roots. Appl Environ Microbiol<br />

58:2691–2693<br />

Sasser M (1990) Identification of bacteria through fatty acid analysis. In: Klement Z,<br />

Rudolph K, Sands D (eds) Methods in phytobacteriology. Akademiai Kiado, Budapest,<br />

pp 199–204<br />

Seghers D, Wittebolle L, Top EM, Verstraete W, Siciliano SD (2004) Impact of agricultural<br />

practices on the Zea mays L. endophytic community. Appl Environ Microbiol 70:1475–<br />

1482<br />

Sessitsch A, Reiter B, Berg G (2004) Endophytic bacterial communities of field-grown potato<br />

plants and their plant growth-promoting abilities. Can J Microbiol 50:239–249<br />

Shishido M, Loeb BM, Chanway CP (1995) External and internal root colonization of lodgepole<br />

pine seedlings by two growth-promoting Bacillus strains originated from different<br />

root microsites. Can J Microbiol 41:707–713


2 Spectrum and Population Dynamics of Bacterial Root Endophytes 31<br />

Smalla K, Wieland G, Buchner A, Zock A, Parzy J, Roskot N, Heuer H, Berg G (2001) Bulk<br />

and rhizosphere soil bacterial communities studied by denaturing gradient gel electrophoresis:<br />

plant dependent enrichment and seasonal shifts. Appl Envrion Microbiol<br />

67:4742–4751<br />

Sturz AV (1995) The role of endophytic bacteria during seed piece decay and potato tuberization.<br />

Plant Soil 175:257–263<br />

Sturz AV, Christie BR, Matheson BG, Nowak J (1997) Biodiversity of endophytic bacteria<br />

which colonize red clover nodules, roots, stems and foliage and their influence on host<br />

growth. Biol Fertil Soils 25:13–19<br />

Sturz AV, Christie BR, Matheson BG, Arsenault WJ, Buchanan NA (1999) Endophytic bacterial<br />

communities in the periderm of potato tubers and their potential to improve<br />

resistance to soil-borne plant pathogens. Plant Pathol 48:360–369<br />

Suckstorff I, Berg G (2003) Evidence of dose-dependent effects on plant growth by<br />

Stenotrophomonas strains from different origins. J Appl Microbiol 95:656–663<br />

Surette MA, Sturz AV, Lada RR, Nowak J (2003) Bacterial endophytes in processing carrots<br />

(Daucus carota L. var. sativus): their localization, population density, biodiversity and<br />

their effects on plant growth. Plant Soil 253:381–390<br />

Trevet IW, Hollis JP (1948) Bacteria in storage organs of healthy plants. Phytopathology<br />

38:960–967<br />

Tsiantos J, Stevens WA (1986) The population dynamics of Corynebacterium michiganense<br />

pv. michiganensis and other selected bacteria in tomato leaves. Phytopathol Mediterr<br />

25:160–162<br />

Van Beneden CA, Keene WE, Strang RA, Werker DH, King AS, Mahon B, Hedberg K,<br />

Bell A, Kelly MT, Balan VK, MacKenzie WR, Fleming D (1999) Multinational outbreak<br />

of Salmonella enterica serotype Newport infections due to contaminated alfalfa sprouts.<br />

JAMA 281:158–162<br />

Vandamme P, Mahenthiralingam E (2003) Strains from the Burkholderia cepacia complex:<br />

Relationship to opportunistic pathogens. J Nematol 35:208–211<br />

Wolf A, Fritze A, Hagemann M, Berg G (2002) Stenotrophomonas rhizophila sp. nov., a novel<br />

plant-associated bacterium with antifungal properties. Int J Evol Syst Microbiol 52:1937–<br />

1944


3<br />

Bacterial Endophytes as Elicitors<br />

of Induced Systemic Resistance<br />

Joseph W. Kloepper, Choong-Min Ryu<br />

3.1<br />

Introduction and Terminology<br />

As indicated elsewhere in this book (e.g. Chap. 1 by Schulz and Boyle),<br />

the question of what are endophytes can be answered in different ways.<br />

For the purposes of this chapter, only those endophytes that could be<br />

isolated from surface-sterilized plant tissue or extracted from within the<br />

plant, as proposed by Hallmann et al. (1997), will be discussed. All of the<br />

rhizobacteria discussed here were isolated by grinding tissues of surfacesterilized<br />

plants, while maintaining sterility controls. It was subsequently<br />

discovered that some of these bacterial strains elicited systemic protection<br />

against pathogens when the bacteria were inoculated onto seeds or into the<br />

potting mix.<br />

Application to crops of many plant-associated bacteria, including some<br />

endophytic bacteria, results in a reduction in the incidence or severity of<br />

diseases. This phenomenon is referred to as biological control. The most<br />

commonly reported mechanism of biological control is antagonism, where<br />

the bacterium causes a reduction in the pathogen population or its diseaseproducing<br />

potential. Antagonism includes the more specific mechanisms<br />

of predation, competition, and antibiosis. Antagonism is discussed in detail<br />

in Chap. 4 by Berg and Hallmann.<br />

An alternative mechanism for biological control is that bacterial metabolites<br />

affect the plant in such a way as to increase the plant’s resistance to<br />

pathogens, a process termed induced systemic resistance (ISR). Resistance<br />

can also be elicited in plants by the application of chemicals or necrosisproducing<br />

pathogens, and this process is termed systemic acquired resistance<br />

(SAR). Pieterse et al. (1998) proposed that ISR and SAR can be differentiatednotonlybytheelicitorbutalsobythesignaltransductionpathways<br />

that are elicited within the plant. Accordingly, ISR is elicited by rhizobac-<br />

Joseph W. Kloepper: Department of Entomology and Plant Pathology, Auburn University,<br />

Auburn, AL 36849, USA, E-mail: kloepjw@auburn.edu<br />

Choong-Min Ryu: Plant Biology Division, The Samuel Roberts Noble Foundation, 2510 Sam<br />

Noble Parkway, Ardmore, OK 73401, USA<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


34 J.W. Kloepper, C.-M. Ryu<br />

teria or other nonpathogenic microorganisms, while SAR is elicited by<br />

pathogens or chemical compounds. Further, the signal transduction pathway<br />

of ISR is independent of salicylic acid but dependent on jasmonate<br />

and ethylene, while the pathway of SAR is dependent on salicylic acid and<br />

shows variable dependency on jasmonate and ethylene. Recent discoveries<br />

show that some rhizobacteria, including some of the endophytic strains<br />

discussed in this chapter, elicit systemic protection that may be dependent<br />

on salicylic acid and independent of jasmonate or ethylene. Hence, ISR<br />

cannot be separated from SAR based on signal transduction pathways. In<br />

this chapter, ISR is used to describe the phenomenon whereby application<br />

of bacteria to one part of the plant results in a significant reduction in the<br />

severity or incidence of a disease following inoculation of a pathogen to<br />

another part of the plant.<br />

3.2<br />

Scope of Endophytes that Elicit Induced Resistance<br />

and Pathosystems Affected<br />

The first indication that endophytic bacteria could elicit ISR dates to 1991<br />

(Wei et al. 1991). Pseudomonas fluorescens strain G8-4, which was later<br />

designated 89B-61 and found to colonize plants internally, elicited systemic<br />

protection against cucumber anthracnose following application to<br />

cucumber seeds.<br />

In efforts to find other strains of endophytic bacteria that elicited ISR,<br />

the research group at Auburn University performed isolations from cucumber<br />

plants in the field or from cucumber seeds. Bacillus pumilus strain<br />

INR7 was isolated from a surface-sterilized stem of a surviving cucumber<br />

plant in a field heavily infested with cucurbit wilt disease, caused by<br />

Erwinia tracheiphila. In two field trials, treatment with INR7 resulted in<br />

significant growth promotion relative to the nontreated control (Wei et al.<br />

1996). In addition, the severity of angular leaf spot, following inoculation<br />

with Pseudomonas syringae pv. lachrymans, and the severity of naturally<br />

occurring anthracnose were significantly reduced by INR7. The cumulative<br />

yield of marketable cucumber fruit was also significantly enhanced by<br />

INR7 in both field trials. In the same study, strain 89B-61 also increased<br />

plant growth and yield and reduced the incidence of both angular leaf spot<br />

and anthracnose. In a subsequent field trial, INR7 reduced the severity of<br />

cucurbit wilt (Zehnder et al. 2001).<br />

Erwinia tracheiphila is completely dependent on the striped cucumber<br />

beetle and the spotted cucumber beetles for survival and transmission. The<br />

finding that cucumber treated with strain INR7 exhibited reduced severity<br />

of cucurbit wilt in the field led to investigations aimed at determining if


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 35<br />

ISR changed beetle feeding activity. In a 2-year field study, the number<br />

of beetles feeding on cucumber was significantly reduced following treatment<br />

with strain INR7 (Zehnder et al. 1997b). In both years of the study,<br />

the season-long average number of cucumber beetles per plant was significantly<br />

lower on plants treated with INR7 than on nonbacterized plants.<br />

ReducedbeetlefeedingonplantstreatedwithstrainINR7wasconfirmedin<br />

subsequent greenhouse studies (Zehnder et al. 1997a). Beetle preference for<br />

nontreated plants was evident within the first 24 h of releasing the beetles<br />

into cages containing cucumber plants. After feeding for 17 days, beetle<br />

damage remained significantly lower on cotyledons and stems of plants<br />

treated with INR7 than on nontreated plants.<br />

Elicitation of altered beetle behavior and feeding preferences in the field<br />

and greenhouse following seed treatment with INR7 was unexpected. As<br />

summarized by Zehnder et al. (1997a), cucumber beetle feeding behavior is<br />

influenced by a group of secondary plant metabolites called cucurbitacins,<br />

which are bitter compounds toxic to most insects. Cucumber beetles consume<br />

cucurbitacins without toxicity, apparently as an evolutionary adaptation<br />

that protects the cucumber beetles from predation. The beetles seek<br />

out cucurbitacins, and concentrations of 1 ng cause cucumber beetles to<br />

demonstrate arrested feeding behavior, whereby the beetles feed intensely<br />

on a single plant without moving from plant to plant. Hence, as an explanation<br />

for reduced beetle feeding on plants treated with INR7, Zehnder et al.<br />

(1997a) reasoned that elicitation of ISR by INR7 might be accompanied by<br />

reduced production of cucurbitacins by cucumber plants. Support for this<br />

hypothesis was found in a study (Zehnder et al. 1997a) in which treatment<br />

of cucumber with strain INR7 resulted in significantly reduced production<br />

of cucurbitacin C. Collectively, the results from studies on cucumber beetles<br />

demonstrate that specific endophytic bacteria can elicit unexpected yet<br />

important physiological changes in plants.<br />

Serratia marcescens strain 90-166 colonizes roots internally (Press et al.<br />

2001) and has been shown to elicit ISR against various diseases of cucumber.<br />

Elicitation of ISR against Fusarium wilt was demonstrated using a splitroot<br />

system (Liu et al. 1995a). Root systems of seedlings were mechanically<br />

separated into two halves, with each half then being placed into a separate<br />

pot. Strain 90-166 was applied to one pot and the pathogen to the other<br />

pot. Numbers of dead plants and severity of the disease were significantly<br />

reduced by strain 90-166 over a 6-week experimental period. In another<br />

study (Liu et al. 1995b), strain 90-166 was found to elicit ISR against angular<br />

leaf spot. Treatment of seeds or cotyledons with strain 90-166 resulted in<br />

significant reductions in numbers and size of angular leaf spot lesions when<br />

the pathogen was inoculated 3 weeks after planting. ISR against angular<br />

leaf spot was also elicited when 90-166 was injected into cotyledons 1 week<br />

before pathogen inoculation. When 90-166 was injected into cotyledons,


36 J.W. Kloepper, C.-M. Ryu<br />

there was a reduction of 1.8 log units in the population of the pathogen<br />

inside leaves.<br />

The capacity of strain 90-166 to elicit protection against cucumber anthracnose<br />

over a 5-week period was evaluated by Liu et al. (1995c). Strain<br />

90-166 was applied to seeds at the time of planting, and Colletotrichum<br />

orbiculare was inoculated onto the first, second, third, fourth, or fifth leaf.<br />

There was approximately 1 week between each leaf stage. Treatment with<br />

strain 90-166 resulted in a significant reduction in the mean total lesion<br />

diameter when the pathogen was inoculated onto the fifth leaf, indicating<br />

that ISR elicited by the strain persists for at least 5 weeks on cucumber.<br />

Elicitation of ISR by strain 90-166 in tobacco against wildfire, caused by<br />

P. syringae pv. tabaci, was also demonstrated by Press et al. (1997). Stem<br />

injection of tobacco with strain 90-166 resulted in a significant decrease<br />

in disease severity when the pathogen was sprayed onto leaves 10 days<br />

after bacterial treatment. Raupach et al. (1996) reported that strain 90-166<br />

also elicited ISR against Cucumber mosaic virus (CMV) on cucumber and<br />

tomato. On cucumber, seed treatment with 90-166 completely prevented<br />

development of CMV symptoms when the virus was inoculated onto cotyledons.<br />

On tomato, the effect of 90-166 was to delay symptom development<br />

over time. The area under the disease progress curve (AUDPC) was significantly<br />

reduced by strain 90-166.<br />

Elicitation of ISR against viruses has also been reported for other strains<br />

of endophytic bacteria. Zehnder et al. (2000) conducted a greenhouse<br />

screen of PGPR (plant growth promoting rhizobacteria) for the potential<br />

to elicit ISR against CMV on tomato. PGPR were applied as seed treatments<br />

and as drenches upon transplanting 2 weeks after seeding. CMV<br />

was rub-inoculated with carborundum onto leaves 1 week after transplanting.<br />

From among 26 tested strains, three strains of endophytes were<br />

selected (Bacillus subtilis strain IN937b, Bacillus pumilus strain SE34, and<br />

Bacillus amyloliquefaciens strain IN937a). All of the selected strains significantly<br />

reduced disease incidence in each of five experiments. In the<br />

same study, Zehnder et al. (2000) conducted two field trials to evaluate<br />

the effects of strains IN937b, SE34, and IN937a on CMV. Treatment with<br />

all three endophytic bacterial strains resulted in significant reductions in<br />

the AUDPC compared to the nonbacterized control in both years of testing.<br />

In another study with CMV in tomato, Murphy et al. (2003) used various<br />

two-strain combinations, where one strain was B. subtilis strain GB03,<br />

which is not reported to be an endophyte, and various endophytic bacteria,<br />

including strains IN937a, IN937b, SE34, and INR7. Spores of the bacteria<br />

were formulated on chitosan as a carrier and this preparation was mixed<br />

into potting mix. All of the bacterial treatments significantly reduced disease<br />

severity based on symptoms, decreased disease incidence based on


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 37<br />

enzyme-linked immunosorbent assay (ELISA), and decreased virus accumulation,<br />

compared to controls.<br />

Three field trials were conducted with various formulations of the endophytic<br />

strains IN937a, IN937b, and SE34 (Murphy et al. 2000) to determine<br />

their capacity to elicit ISR against Tomato mottle virus (ToMoV), which is<br />

vectored by the silver whitefly (Bremisia argentifolii). The plots were inoculated<br />

with ToMoV by natural movement of viruliferous whitefly adults<br />

from adjacent plantings of ToMoV-resistant tomato germplasm that was<br />

inoculated prior to transplanting into the field. The incidence and severity<br />

of ToMoV were significantly reduced by one or more of the formulations<br />

of each endophyte. The number of whitefly nymphs detected on plants was<br />

significantly reduced by strains IN937a and IN937b. Hence, as in the case<br />

with cucurbit wilt of cucumber, some endophytic bacteria can elicit ISR<br />

against both a plant disease and its insect vectors.<br />

ISR elicited by the endophytes B. pumilus strain SE34, S. marcescens strain<br />

90-166, and Pseudomonas fluorescens strain 89B-61 has been shown to reduce<br />

the severity of blue mold of tobacco, caused by Peronospora tabacina<br />

(Zhang et al. 2002a, 2002b, 2004). In one study (Zhang et al. 2002b), strains<br />

SE34, 90-166, and 89B-61 elicited ISR in detached leaf and microtiter plate<br />

bioassays as well as in pot trials in the greenhouse. In the pot assay, application<br />

of all three strains as a soil drench to 4-week-old plants of three<br />

tobacco cultivars resulted in significant reductions in the mean percentage<br />

of leaf area with lesions caused by P. tabacina inoculated onto leaves<br />

1 week after bacterial treatment. Sporulation of the pathogen on lesions<br />

was significantly decreased by treatment with the three strains in pot trials.<br />

Strains SE34, 90-166, and 89B-61 also significantly reduced disease severity<br />

in the detached leaf (injection of a bacterial suspension into petioles) and<br />

microtiter plate bioassays (application of bacterial suspensions to roots).<br />

Sporulation of the pathogen was significantly reduced by both strains in<br />

the detached leaf bioassay.<br />

In another study (Zhang et al. 2004), strains SE34 and 90-166 were used<br />

to explore the relationship between elicitation of plant growth promotion<br />

and ISR. Application of the endophytes as a seed treatment alone elicited<br />

significantly enhanced tobacco plant growth but not disease protection.<br />

When the strains were applied as seed treatments followed by a soil drench,<br />

both plant growth promotion and ISR were elicited. Overall, the results from<br />

this study indicated that while there was a relationship between growth<br />

promotion and ISR, elicitation of growth promotion can occur without<br />

elicitation of ISR; however, when ISR was elicited, growth promotion was<br />

also elicited with the bacterial strains used in the study.<br />

Endophytic bacteria have also elicited ISR against tomato late blight,<br />

caused by Phytophthora infestans (Yan et al. 2002). Application of strains<br />

SE34 and 89B-61 by incorporation into the potting medium at the time of


38 J.W. Kloepper, C.-M. Ryu<br />

planting elicited significant reductions in disease severity when P. infestans<br />

wasinoculatedontoleaves5weeksafterplanting.<br />

Greenhouse screening of endophytic Bacillus spp. that have elicited ISR<br />

on some crops was conducted in Thailand (Jetiyanon and Kloepper 2002)<br />

as a first step toward employment of endophyte-elicited ISR in tropical<br />

agriculture. This study used four different host/pathogen systems: tomato<br />

and Ralstonia solanacearum, long cayenne pepper (Capsicum annuum<br />

var. acuminatum) andColletotrichum gloeosporioides, greenkuangfutsoi<br />

(Brassica chinensis var. parachinensis)andRhizoctonia solani, and cucumber<br />

and CMV. The goal of the study was to find mixtures of endophytic<br />

spore-forming bacteria that elicited ISR in all four host/pathogen systems.<br />

Seven individual strains and 11 combinations of 2 strains were tested. One<br />

strain (B. amyloliquefaciens IN937a) and four mixtures (IN937a + B. subtilis<br />

IN937b; IN937b + B. pumilus SE34; IN937b + B. pumilus SE49; and IN937b<br />

+ B. pumilus INR7) significantly reduced incidence or severity of all four<br />

diseases. The results are noteworthy for two reasons. First, they indicate<br />

that ISR elicited by specific endophytic bacterial strains can protect hosts<br />

under tropical conditions. Second, the results show that mixtures of two<br />

bacterial strains are superior to individual strains for eliciting significant<br />

protection in multiple hosts against different pathogens.<br />

FurtherevidencefortheconceptofusingstrainmixturesofBacillus spp.<br />

to increase the repeatability of plant growth promotion or elicitation of ISR<br />

by bacteria was reported in a follow-up field investigation (Jetiyanon et al.<br />

2003). Field tests were conducted in Thailand to find mixtures of bacteria<br />

that could protect several different hosts against the multiple diseases<br />

that are typical under the multi- or inter-cropping agricultural conditions<br />

predominant in Thailand. In tests conducted during the rainy and dry<br />

seasons, some two-strain mixtures of endophytes more consistently protected<br />

against disease than did a single strain. In each season, the mixture<br />

of strains IN937a and IN937b significantly protected against all the tested<br />

diseases (southern blight of tomato, CMV on cucumber, and anthracnose<br />

of long cayenne pepper). The same mixture of endophytes also resulted in<br />

significant yield increases of all crops during the rainy season.<br />

Endophytic bacteria have also been shown to elicit ISR in conifers.<br />

Enebak and Carey (2000) tested the potential elicitation of ISR on loblolly<br />

pine (Pinus taeda) bystrainsB. sphaericus SE56 and B. pumilus strains<br />

INR7, SE34, SE49, and SE52 against Cronartium quercuum f. sp. fusiforme,<br />

which causes fusiform rust. Bacteria were applied at seeding, and suspensions<br />

of C. quercuum basidiospores from field-collected telia on water<br />

oak (Quercus nigra) were sprayed onto the pine seedlings at five different<br />

times. Six months after the final application of basidiospores, the incidence<br />

of fusiform rust was determined by noting the presence or absence of the<br />

typical symptoms of main-stem swellings or galls. The experiment was


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 39<br />

conducted annually for 2 years. All strains except SE49 resulted in significant<br />

reductions in disease incidence in either one of the 2 years or in the<br />

pooled data from both years.<br />

Before concluding this discussion of case studies of endophytic bacteria<br />

that have been shown to elicit ISR, a note should be made about Azospirillum<br />

spp. Azospirillum brasilense is a well characterized endophyte, and nearly<br />

all strains of this species have been shown to promote growth of many<br />

crop species (Bashan and de-Bashan 2002). Because many of the strains of<br />

Bacillus spp. cited above that elicit ISR also elicit plant growth promotion,<br />

one might expect that A. brasilense would also elicit ISR. However, this<br />

is not the case. Bashan and de-Bashan (2002) investigated the potential<br />

of A. brasilense to elicit ISR in tomato against bacterial speck, caused by<br />

P. syringae pv. Tomato, and concluded that this endophyte does not elicit<br />

ISR.<br />

3.3<br />

Internal Colonization of Endophytes<br />

that Elicit Induced Resistance<br />

In most of the studies discussed in the previous section, extensive microbial<br />

ecology studies to determine the extent of internal colonization of plant<br />

tissues by the applied endophytes were not carried out. Typically, isolations<br />

are performed near the location where the pathogen was applied. Such<br />

isolation is done to test one of the suggested tenants of ISR: that there<br />

is physical separation of the pathogen and the inducing agent. According<br />

to this tenant, physical separation is required to differentiate ISR from<br />

antagonism as a mechanism for protection against pathogens. While testing<br />

for physical separation has validity, it also creates an inherent problem<br />

with endophytic bacteria that exhibit systemic colonization of plants. An<br />

endophyte that can move within the plant and colonize petioles could,<br />

theoretically, still elicit ISR at a level sufficient to reduce disease severity<br />

ofafoliarpathogen.However,basedonthetenantofphysicalseparation,<br />

one could not state that ISR was the operable mechanism by which disease<br />

severity was reduced. Hence, some endophytic bacteria might actually elicit<br />

plant defense although they are not spatially separated from the pathogen.<br />

An example illustrating the limited internal colonization of well characterized<br />

endophytes that elicit ISR is P. fluorescens strain 89B-61, which<br />

was initially designated as strain G8-4. Because this strain has reactions<br />

in biochemical tests that are intermediate between P. putida and P. fluorescens,<br />

some publications before 1997 refer to 89B-61 as P. putida.Later<br />

publications designate the strain as P. fluorescens basedonrepeatedfatty<br />

acid analyses. Chen et al. (1995) found that strain 89B-61 significantly


40 J.W. Kloepper, C.-M. Ryu<br />

reduced the severity of Fusarium wilt of cotton. In this system, 89B-61<br />

was stab-inoculated into seedling stems 13 days prior to inoculation with<br />

Fusarium oxysporum f. sp. vasinfectum atapoint1.5cmabovethepoint<br />

of bacterial inoculation. Using a rifampicin-resistant mutant of 89B-61, no<br />

movement up the stem from the point of bacterial inoculation was detected.<br />

In a separate greenhouse study, Kloepper et al. (1992) reported that seed<br />

treatment of cucumber with strain G8-4 (89B-61) significantly reduced lesion<br />

numbers and size of anthracnose following challenge inoculation of<br />

leaves with C. orbiculare. This induced resistance was associated with colonization<br />

of internal root tissues at log 4.0 cfu/g at 14 days after emergence;<br />

however, the bacteria were not isolated from stems or leaves. Elicitation of<br />

induced resistance in cucumber by 89B-61 was confirmed in field studies<br />

(Wei et al. 1996), where application of the bacterium as a seed treatment<br />

resulted in significant reductions in severity of anthracnose and angular<br />

leaf spot.<br />

Quadt-Hallmann et al. (1997) used isolation, ELISA, and immunogold<br />

labeling with P. fluorescens 89B-61-specific polyclonal antibodies to investigate<br />

the pattern of internal colonization of cotton by the ISR-eliciting<br />

strain 89B-61. Results from isolation studies indicated that, following treatment<br />

of seeds, 89B-61 colonized roots internally at a mean population of<br />

1.1×10 3 cfu/g, while the bacterium was not recovered from stems, cotyledons,<br />

or leaves. With ELISA, strain 89B-61 was detected outside and inside<br />

roots but not inside stems, cotyledons, or leaves. Electron microscopy with<br />

immunogold labeling revealed that internal colonization of roots by 89B-61<br />

was restricted mainly to intercellular spaces of the epidermis. Interestingly,<br />

the colonization pattern was quite distinct from that of another bacterium<br />

(Enterobacter asburiae strain JM22) that does not elicit ISR. JM22 colonized<br />

throughout the root cortex, including inside the vascular stele, in intercellular<br />

spaces close to the conducting elements, as was previously found in<br />

other plant species (Quadt-Hallmann and Kloepper 1996). It was suggested<br />

that the internal colonization of cotton by 89B-61 and JM22 could be considered<br />

as representative of two fundamental options for how endophytic<br />

bacteria colonize plants after application to seeds or soils. The first pattern<br />

is that of 89B-61 and consists of limited internal colonization of roots. The<br />

second pattern, demonstrated by JM22, consists of extensive internal root<br />

colonization and ultimately in some vascular colonization.<br />

Internal colonization of roots by S. marcescens strain 90-166 was demonstrated<br />

by Press et al. (2001) in an investigation into the role of iron in<br />

elicitation of ISR. Cucumber root colonization by the wild-type strain was<br />

compared to that by a mutant deficient in siderophore production. While<br />

the total root population sizes (external and internal colonization) were statistically<br />

equivalent, the internal population size was significantly greater<br />

with the wild-type than with the siderophore-negative mutant. Because the


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 41<br />

mutant failed to elicit ISR against anthracnose, while ISR was elicited by<br />

the wild-type strain, it was concluded that capacity to elicit ISR was related<br />

to the population size of the bacterium inside roots.<br />

3.4<br />

Plant Responses to Endophytic Elicitors<br />

Investigations aimed at determining how plants respond to inoculation<br />

with endophytes that elicit ISR is one approach to studying mechanisms<br />

of ISR by such bacteria. During the previously discussed study on cotton<br />

colonization by strain 89B-61 (Quadt-Hallmann et al. 1997), the epidermal<br />

cell walls of plant cells adjacent to cells of 89B-61 in the intercellular space<br />

developed electron-opaque appositions of an amorphous matrix.<br />

Two cytological studies were conducted by Benhamou et al. (1996, 1998)<br />

with B. pumilus strain SE34. In the first study (Benhamou et al. 1996),<br />

colonization of pea roots by Fusarium oxysporum f. sp. pisi was restricted<br />

totheepidermisandoutercortexofrootstreatedwithSE34,whilein<br />

nonbacterized roots, the pathogen colonized the cortex, endodermis, and<br />

the paratracheal parenchyma cells. This reduction in fungal colonization by<br />

SE34wasassociatedwithstrengtheningoftheepidermalandcorticalcell<br />

walls. In addition, roots treated with SE34 exhibited newly formed barriers<br />

beyondthesiteoffungalinfection.Thesebarrierswerecellwallappositions<br />

that contained large amounts of callose and were infiltrated with phenolic<br />

compounds. Phenolic compounds were detected in transmission electron<br />

microscopy using gold-complexed laccase and were found to accumulate<br />

in host cell walls, in intercellular spaces, and on the surface of and inside<br />

the invading pathogen hyphae.<br />

In another study (Benhamou et al. 1998), the effect of SE34 alone or in<br />

combination with chitin on structural and cytochemical changes of tomato<br />

infected with F. oxysporum f. sp. radicis-lycopersici was investigated. Treatment<br />

with SE34 reduced the severity of typical symptoms, including wilting<br />

of seedlings and numbers of brown lesions on lateral roots. This disease<br />

protection by strain SE34 was associated with more limited fungal colonization<br />

of roots and with marked changes in host physiology. Physiological<br />

changes elicited by strain SE34 included an increase in host cell wall density,<br />

the accumulation of polymorphic deposits at sites of potential pathogen<br />

penetration, and the occlusion of epidermal cells and intercellular spaces<br />

with an osmophilic, amorphous material that appeared to trap the invading<br />

fungal hyphae. The extent and magnitude of the physiological changes in<br />

the host elicited by SE34 were enhanced by the addition of chitosan. Interestingly,<br />

the overall chitin component of the pathogen was structurally<br />

preserved in roots treated with SE34 with or without chitosan at the time


42 J.W. Kloepper, C.-M. Ryu<br />

when hyphal degradation was apparent. This suggests that synthesis of<br />

chitinase in bacteria-treated roots is not an early event in the cascade of<br />

physiological steps in signal transduction that lead to induced resistance.<br />

Benhamou et al. (1998) concluded: “According to our cytological observations,<br />

the induction of resistance triggered by B. pumilus strain SE34<br />

involves a sequence of events including first the elaboration of structural<br />

barriers and the production of toxic substances such as phenolics and phytoalexins,<br />

and second the synthesis and accumulation of other molecules<br />

including chitinases and other hydrolytic enzymes such asβ-1,3-glucanases<br />

which probably contribute to the release of oligosaccharides that, in turn,<br />

can stimulate other defense reactions.”<br />

Jeun et al. (2004) conducted a cytological comparison of cucumber plants<br />

in which systemic resistance had been elicited by bacteria or by chemicals.<br />

In this study, the endophytes 89B-61 and 90-166 were used to elicit ISR<br />

against C. orbiculare. Significantly fewer numbers of anthracnose lesions<br />

developed on plants treated with 89B-61 and 90-166 than on the control<br />

(chemical treatment). Cytological studies using fluorescent microscopy revealed<br />

a higher frequency of autofluorescent epidermal cells, which are<br />

related to accumulation of phenolic compounds, at the sites of fungal penetration<br />

in plants treated with either strain and inoculated with C. orbiculare.<br />

In addition, callose-like structures (β-1,3-glucan polymers) were frequently<br />

deposited at the site of fungal penetration of the leaves of plants treated<br />

with either strain.<br />

Investigations on plant responses to elicitation of ISR can also examine<br />

the signal transduction pathway of plants to determine general biochemical<br />

pathways in plants during ISR. As previously discussed, according to the<br />

model pathway for signal transduction (Pieterse et al. 1998), ISR pathways<br />

are independent of salicylic acid, but dependent on ethylene, jasmonic acid,<br />

and the regulatory gene npr-1. Further, according to the model, ISR elicited<br />

by bacteria does not result in the accumulation of pathogenesis-related (PR)<br />

proteins. PR proteins are accumulated during SAR elicited by pathogens<br />

and chemicals, and SAR is dependent upon salicylic acid.<br />

Afewstudiesonsignalpathwayshavebeenreportedwithendophytic<br />

bacteria that elicit ISR. In the tomato late blight system, ISR was elicited<br />

by B. pumilus strain SE34 on NahG lines, which breakdown endogenous<br />

salicylic acid, but not in the ethylene-insensitive NR/NR line or in the<br />

jasmonic acid-insensitive df1/df1 line (Yan et al. 2002). These results are<br />

consistent with the model of Pieterse et al. (1998). Similar results were<br />

reported by Zhang et al. (2002a). In the tobacco blue mold system, B. pumilus<br />

strain SE34, as well as two strains of Gram-negative bacteria, elicited ISR<br />

on both wild-type and NahG transgenic tobacco lines, as evidenced by<br />

significant reductions in the severity of blue mold on bacterized plants<br />

compared to nonbacterized plants.


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 43<br />

Different results were found with strain 89B-61 (Park and Kloepper<br />

2000), which elicits ISR in tobacco against wildfire caused by P. syringae pv.<br />

tabaci. In this system, a transgenic line of tobacco with a β-glucuronidase<br />

(GUS) reporter gene fused to the PR-1a promoter had significantly reduced<br />

severity of wildfire compared to nonbacterized controls. Elicitation of ISR<br />

by strain 89B-61 was associated with a significant increase in GUS activity<br />

in microtiter plate and whole plant bioassays. Hence, with strain 89B-61,<br />

elicitation of ISR results in activation of the PR-1a gene, which is activated<br />

during SAR but not during bacterial-induced ISR according to the model<br />

of Pieterse et al. (1998).<br />

Signal pathways in ISR elicited by P. fluorescens strain CHA0 in Arabidopsis<br />

against Peronospora parasitica were investigated by Iavicoli et al.<br />

(2003) using various transgenic and mutant plant lines: NahG (for degradation<br />

of salicylic acid), sid2-1 (lacks production of salicylic acid), npr1-1<br />

(nonexpressor of PR genes), jar1-1 (insensitive to jasmonic acid), ein2-1<br />

(insensitive to ethylene), eir1-1 (insensitive to ethylene/auxin), and pad2-1<br />

(phytoalexin-deficient). ISR was elicited by strain CHA0 in all lines except<br />

jar1-1, eir1-1,andnpr1-1.<br />

In another study of signaling pathways, Ryu et al. (2003b) used endophytic<br />

bacterial strains in Arabidopsis to elicit ISR against two different<br />

pathovars of P. syringae (pvs. tomato and maculicola). Strains SE34, 90-<br />

166, and 89B-61 elicited ISR against both pathogens. Strain SE34 elicited<br />

a salicylic acid-independent pathway against one pathovar and salicylic<br />

acid-dependent pathway against a different pathovar. Additional tests of<br />

strains 89B-61 and SE34 on various mutant lines of Arabidopsis (Ryu et al.<br />

2003b) revealed that, in agreement with the model of Pieterse et al. (1998),<br />

ISRelicitedbybothstrainswasdependentonNPR1andISRelicitedby<br />

SE34 was dependent on jasmonic acid and ethylene. However, in contrast<br />

to the model, ISR elicited by strain 89B-61 was independent of ethylene and<br />

jasmonic acid, and ISR by strain 90-166 was dependent on jasmonic acid<br />

but independent of ethylene.<br />

Endophyte strains 90-166 and SE34 also elicited ISR against CMV in<br />

Arabidopsis (Ryu et al. 2004b). Strains 90-166 and SE34 reduced disease<br />

severity in NahG plants, indicating that ISR elicited by strains 90-166 and<br />

SE34 was independent of salicylic acid. Further investigation on the signal<br />

pathway of ISR against CMV elicited by strain 90-166 with lines NahG,<br />

npr1, andfad3-2 fad7-2 fad8 (insensitive to jasmonic acid) indicated that<br />

ISR against CMV by strain 90-166 is independent of salicylic acid and NPR1,<br />

but is dependent on jasmonic acid (Ryu et al. 2004b).<br />

Collectively, the results on signaling pathways of ISR elicited by endophytic<br />

bacteria indicate that different pathways are elicited by various<br />

strains. Further, the specific signal transduction pathway that is activated<br />

during ISR depends on the host plant and, at least in one case, on the<br />

pathogen used on a given host.


44 J.W. Kloepper, C.-M. Ryu<br />

A new approach to investigations on plant responses to endophytic bacteriawasrecentlyopenedbythefindingthatvolatileorganiccompounds<br />

produced by the endophyte B. amyloliquefaciens IN937a elicit plant growth<br />

promotion (Ryu et al. 2003a) and ISR (Ryu et al. 2004a). Significant growth<br />

promotion of Arabidopsis by IN937a was observed in I-plates, which have<br />

a raised plastic divider separating agar on each half of the dish, thus preventing<br />

movement of soluble compounds. When IN937a was placed on<br />

onesideofanI-plate,Arabidopsis plants growing on the other side exhibited<br />

enhanced growth, presumably as a result of volatiles produced by<br />

the bacteria. Characterization of the volatile organic compounds (VOCs)<br />

produced by IN937a, coupled with bioassays of fractions of VOCs, revealed<br />

that 2,3-butanediol and acetoin elicited plant growth promotion. In a separate<br />

study (Ryu et al. 2004a), exposure of Arabidopsis to VOCs from strain<br />

IN937a resulted in significantly less disease caused by Erwinia carotovora<br />

subsp. carotovora. Tests with various mutant lines of Arabidopsis revealed<br />

that elicitation of ISR by VOCs of IN937a is independent of jasmonic acid,<br />

ethylene, salicylic acid, and npr1. Such a pattern of signal pathway has not<br />

been reported with ISR elicited by bacteria and, therefore, it is likely that<br />

VOCs of IN937a elicit a distinct, and as yet uncharacterized, pathway in<br />

Arabidopsis.<br />

3.5<br />

Implementation in Production Agriculture:<br />

Two Case Studies<br />

The principle that endophytic bacteria can elicit ISR or plant growth promotion<br />

has been extended to use in production agriculture and horticulture<br />

through the development of two products. These two products are discussed,<br />

not as endorsements of the products, but as case studies indicating<br />

that our growing scientific knowledge of endophytic bacteria can be put to<br />

practical use.<br />

In the first case study, an agricultural product has been developed using<br />

the capacity of a single endophytic strain of Bacillus spp. to elicit<br />

both ISR and plant growth promotion. The product is Yield Shield, which<br />

is produced by Gustafson, LLC. Yield Shield consists of a spore preparation<br />

of the B. pumilus strain listed in this review as INR7 (Table 3.1)<br />

and designated by Gustafson as GB34 (http://www.gustafson.com/Labels/<br />

yield shield label.pdf). The product received registration from the United<br />

States Enivironmental Protection Agency (EPA) in 2003 for use on soybeans<br />

to protect against Rhizoctonia solani and Fusarium spp. Seed treatment of<br />

soybean with Yield Shield and strain INR7 results in significant seedling<br />

growth promotion and in ISR, which is apparent both by a significant de-


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 45<br />

Table 3.1. Endophytic bacterial strains that have been reported to elicit induced systemic<br />

resistance (ISR) in at least two publications<br />

Strain no. and<br />

identification<br />

IN937a<br />

Bacillus amyloliquefaciens<br />

IN937b<br />

Bacillus<br />

subtilis<br />

90-166<br />

Serratia<br />

marcescens<br />

Effectsreportedandsystemsused a Reference<br />

Reduced incidence or severity of vegetable diseases<br />

(caused by Cucumber mosaic virus (CMV), Sclerotium<br />

rolfsii, Ralstonia solanacearum, Colletotrichum<br />

gloeosporioides,andRhizoctonia solani) in greenhouse<br />

and field trials in Thailand<br />

Jetiyanon et al.<br />

2002;<br />

Jetiyanon and<br />

Kloepper 2003<br />

Reduced incidence of CMV on tomato in the greenhouse Zehnder et al.<br />

2000<br />

When applied to tomato with the non-endophyte Bacillus<br />

subtilis GB03, reduced severity of CMV in the greenhouse<br />

Reduced incidence and severity of tomato mottle virus in<br />

the field. Also reduced numbers of the white fly vector<br />

feeding on plants<br />

Volatile organic compounds of the strain elicit growth<br />

promotion of Arabidopsis and ISR against Erwinia<br />

carotovora subsp. carotovora<br />

Component of the product, BioYield (Gustafson LLC,<br />

http://www.gustafson.com)<br />

Murphy et al.<br />

2003<br />

Murphy et al.<br />

2000<br />

Ryu et al.<br />

2003a, 2004<br />

Kloepper et al.<br />

2004<br />

Reduced incidence of CMV on tomato in the greenhouse Zehnder et al.<br />

2000<br />

When applied to tomato with the non-endophyte<br />

B. subtilis GB03, reduced severity of CMV in the<br />

greenhouse<br />

Reduced incidence and severity of tomato mottle virus in<br />

the field. Also reduced numbers of the white fly vector<br />

feeding on plants<br />

Reduced incidence and delayed development of<br />

symptoms of Fusarium wilt of cucumber, caused by<br />

Fusarium oxysporum f. sp. cucumerinum<br />

Reduced severity of bacterial angular leaf spot of<br />

cucumber, caused by Pseudomonas syringae pv.<br />

lachrymans<br />

Reduced severity of anthracnose, caused by<br />

Colletotrichum orbiculare, in two cucumber cultivars<br />

over a 5-week period after bacterial treatment<br />

Reduced incidence of CMV on cucumber and tomato<br />

and reduced the area under the disease progress curve<br />

(AUDPC)<br />

Decreased severity of tobacco wildfire, caused by<br />

Pseudomonas syringae pv. tabaci<br />

Murphy et al.<br />

2003<br />

Murphy et al.<br />

2000<br />

Liu et al. 1995a<br />

Liu et al. 1995b<br />

Liu et al. 1995c<br />

Raupach et al.<br />

1996<br />

Press et al.<br />

1997


46 J.W. Kloepper, C.-M. Ryu<br />

Table 3.1. (continued)<br />

Strain no. and<br />

identification<br />

SE34 Bacillus<br />

pumilus<br />

Effectsreportedandsystemsused a Reference<br />

Wild-type strain reduced severity of cucumber<br />

anthracnose. A siderophore-negative mutant did not elicit<br />

ISR and colonized roots internally at lower populations<br />

than the wild-type<br />

Decreased severity of tobacco blue mold, caused by<br />

Peronospora tabacina, in NahG transgenic tobacco lines<br />

that degrade salicylic acid<br />

Decreased severity of tobacco blue mold in microtiter<br />

plate assays and detached leaf assays. Reduced pathogen<br />

sporulation<br />

Reduced symptoms of Pseudomonas syringae pvs. tomato<br />

and maculicola on Arabidopsis<br />

Decreased severity of tobacco blue mold, caused by<br />

Peronospora tabacina<br />

Reduced incidence or severity of vegetable diseases<br />

(caused by cucumber mosaic virus, Sclerotium rolfsii,<br />

Ralstonia solanacearum, Colletotrichum gloeosporioides,<br />

and Rhizoctonia solani) in greenhouse and field trials in<br />

Thailand<br />

Decreased severity of tobacco blue mold, caused by<br />

Peronospora tabacina, in NahG transgenic tobacco lines<br />

that degrade salicylic acid<br />

Decreased severity of tobacco blue mold in microtiter<br />

plate assays and detached leaf assays. Reduced pathogen<br />

sporulation<br />

Decreased severity of tobacco blue mold when applied as<br />

both seed treatment and drench in the greenhouse.<br />

Elicitation of ISR was associated with growth promotion<br />

Decreased severity of tomato late blight, caused by<br />

Phytophthora infestans, and decreased germination of<br />

sporangia and zoospores of the pathogen<br />

Press et al.<br />

2001<br />

Zhang et al.<br />

2002a<br />

Zhang et al.<br />

2002b<br />

Ryu et al.<br />

2003b<br />

Zhang et al.<br />

2004<br />

Jetiyanon et.<br />

al. 2002;<br />

Jetiyanon and<br />

Kloepper 2003<br />

Zhang et al.<br />

2002a<br />

Zhang et al.<br />

2002b<br />

Zhang et al.<br />

2004<br />

Yan et al. 2002<br />

Reduced incidence of CMV on tomato in the greenhouse Zehnder et al.<br />

2000<br />

When applied to tomato with the non-endophyte<br />

B. subtilis GB03, reduced severity of CMV in the<br />

greenhouse<br />

Reduced incidence and severity of tomato mottle virus in<br />

the field<br />

Reduced incidence of Fusiform rust, caused by<br />

Cronartium quercuum f. sp. fusiforme,onloblollypine<br />

Murphy et al.<br />

2003<br />

Murphy et al.<br />

2000<br />

Enebak and<br />

Carey 2000


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 47<br />

Table 3.1. (continued)<br />

Strain no. and<br />

identification<br />

INR7<br />

Bacillus<br />

pumilus<br />

(available as<br />

the product<br />

Yield Shield;<br />

Gustafson<br />

LLC)<br />

89B-61<br />

Pseudomonas<br />

fluorescens<br />

(earlier<br />

referred to<br />

as G8-4)<br />

Effectsreportedandsystemsused a Reference<br />

Restricted colonization of pea roots by F. oxysporum f. sp.<br />

pisi and induced formation of structural barriers in the<br />

plant<br />

Reduced damage of tomato roots to F. oxysporum f. sp.<br />

radicis-lycopersici and induced structural and<br />

cytochemical barriers in the plant<br />

Reduced symptoms of Pseudomonas syringae pvs. tomato<br />

and maculicola on Arabidopsis<br />

Reduced incidence or severity of vegetable diseases<br />

(caused by CMV, Sclerotium rolfsii, Ralstonia<br />

solanacearum, Colletotrichum gloeosporioides,and<br />

Rhizoctonia solani) in greenhouse and field trials in<br />

Thailand<br />

Decreased severity of anthracnose (caused by<br />

Colletotrichum orbiculare) and angular leaf spot (caused<br />

by Pseudomonas syringae pv. lachrymans) on cucumber<br />

in field trials<br />

Decreased the incidence of cucurbit wilt disease, caused<br />

by Erwinia tracheiphila,infieldtrials<br />

Decreased numbers of cucumber beetles on plants in the<br />

field<br />

Decreased beetle feeding activity and transmission of<br />

E. tracheiphila on cucumber in cages where beetles had<br />

a choice between bacterial-treated and nontreated plants<br />

When applied to tomato with the non-endophyte<br />

B. subtilis GB03, reduced severity of CMV in the<br />

greenhouse<br />

Reduced incidence of Fusiform rust, caused by<br />

Cronartium quercuum f. sp. fusiforme,onloblollypine<br />

Decreased severity of anthracnose (caused by<br />

Colletotrichum orbiculare) and angular leaf spot (caused<br />

by Pseudomonas syringae pv. lachrymans) on cucumber<br />

in field trials<br />

Decreased severity of tobacco blue mold, caused by<br />

Peronospora tabacina<br />

Decreased severity of Fusarium wilt of cotton, caused by<br />

Fusarium oxysporum f. sp. vasinfectum<br />

Following seed treatment of cotton, 89B-61 reached an<br />

internal root population of 1.1×10 3 cfu/g. Bacteria were<br />

not detected inside cotyledons, stems, or leaves<br />

Benhamou et<br />

al. 1996<br />

Benhamou et<br />

al. 1998<br />

Ryu et al.<br />

2003b<br />

Jetiyanon et.<br />

al. 2002;<br />

Jetiyanon and<br />

Kloepper 2003<br />

Wei et al. 1996<br />

Zehnder et al.<br />

2001<br />

Zehnder et al.<br />

1997b<br />

Zehnder et al.<br />

1997a<br />

Murphy et al.<br />

2003<br />

Enebak and<br />

Carey 2000<br />

Wei et al. 1996<br />

Zhang et al.<br />

2004<br />

Chen et al.<br />

1995<br />

Quadt-<br />

Hallmann et<br />

al. 1997


48 J.W. Kloepper, C.-M. Ryu<br />

Table 3.1. (continued)<br />

Strain no. and<br />

identification<br />

CHA0<br />

Pseudomonas<br />

fluorescens<br />

Effectsreportedandsystemsused a Reference<br />

Decreased severity of anthracnose (caused by<br />

Colletotrichum orbiculare) and angular leaf spot (caused<br />

by Pseudomonas syringae pv. lachrymans) on cucumber<br />

Decreased severity of tomato late blight, caused by<br />

Phytophthora infestans, and decreased germination of<br />

sporangia and zoospores of the pathogen when applied to<br />

seeds<br />

Decreased severity of tobacco blue mold, caused by<br />

Peronospora tabacina, in NahG transgenic tobacco lines<br />

that degrade salicylic acid<br />

Decreased severity of tobacco blue mold in microtiter<br />

plate assays and detached leaf assays. Reduced pathogen<br />

sporulation<br />

Decreased severity of tobacco wildfire, caused by<br />

Pseudomonas syringae pv. tabaci. Activated the promoter<br />

for PR1a [a pathogenesis-related (PR) protein]<br />

Reduced symptoms of Pseudomonas syringae pvs. tomato<br />

and maculicola on Arabidopsis<br />

Reduced number of lesions of Colletotrichum orbiculare<br />

on cucumber and increased deposition of callose-like<br />

polymers on leaf cells at the site of pathogen penetration<br />

Reduced mean numbers and size of anthracnose lesions<br />

on cucumber<br />

Colonized roots internally at log 4.0 cfu/g at 2 weeks after<br />

emergence when applied as seed treatments. Bacteria<br />

were not detected in leaves or stems<br />

Reduced severity of Tobacco necrosis virus (TNV) on<br />

tobacco.<br />

Severity of TNV on tobacco was reduced equivalently by<br />

the wild-type strain and a transgenic strain carrying<br />

pchAB genes for synthesis of salicylic acid.<br />

Reduced sporulation of Peronospora parasitica on<br />

Arabidopsis<br />

Wei et al. 1996<br />

Yan et al. 2002<br />

Zhang et al.<br />

2002a<br />

Zhang et al.<br />

2002b<br />

Park and<br />

Kloepper 2000<br />

Ryu et al.<br />

2003b<br />

Jeun et al. 2004<br />

Wei et al. 1991<br />

Kloepper et al.<br />

1992<br />

Maurhofer et<br />

al. 1994<br />

Maurhofer et<br />

al. 1998<br />

Iavicoli et al.<br />

2003<br />

a In all cases, the stated reductions in disease incidence or severity and the effects on insects<br />

are statistically significant at P ≤ 0.05<br />

crease in incidence and severity of R. solani inoculated onto stems at a point<br />

where INR7 does not colonize, and by a systemic increase in lignification<br />

of plant cell walls (C.-M. Ryu and C.-H. Hu, unpublished). It should be<br />

emphasized that Yield Shield is a unique case for a rhizobacterium that


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 49<br />

elicits ISR in that economically significant efficacy sufficient to warrant the<br />

costs of product development and EPA registration was shown for a single<br />

bacterial strain.<br />

In the second case study, the product consists of a two-strain mixture<br />

of Bacillus spp., where one strain (IN937a) is an endophyte that elicits ISR<br />

(Table 3.1) and the other strain (GB03) is a non-endophyte. The product is<br />

BioYield and is also produced by Gustafson. The development of BioYield<br />

was recently reviewed (Kloepper et al. 2004). The underlying concept was<br />

to develop a biological formulation consisting of components known to exert<br />

different mechanisms for control of diseases. The selected components<br />

and their mechanisms were chitosan (as a carrier) for nematode control<br />

via promotion of indigenous soil predators and antagonists to root-knot<br />

nematodes, B. subtilis strain GB03 for control of soil-borne pathogens via<br />

production of the antibiotic iturin, and one of several tested endophytic<br />

Bacillus spp. that elicit ISR. The most unexpected finding was that the<br />

three-component combination (chitosan plus two bacterial strains) exhibited<br />

more consistent, and a greater magnitude of, growth promotion<br />

and systemic protection against pathogens than did any of the individual<br />

components (Kloepper et al. 2004). Based on the results, the two-strain<br />

combination of B. amyloliquefaciens strain IN937a and B. subtilis strain<br />

GB03 was selected for product development.<br />

3.6<br />

Conclusions<br />

As discussed in this review, selected strains of nonpathogenic endophytic<br />

bacteria can elicit ISR in plants, leading to reductions in severity of various<br />

diseases. Research on such endophytes has concentrated both on delineating<br />

the pathosystems where protection results and in understanding<br />

plant responses that occur during the signal transduction pathways that<br />

culminate in disease protection. In many cases, elicitation of ISR by endophytic<br />

bacilli is associated with increased plant growth, and the relationship<br />

between ISR and growth promotion should be further investigated. Elucidation<br />

of specific bacterial determinants that account for elicitation of<br />

ISR is just beginning, and further work is needed to understand why one<br />

strain of a given bacterial species can elicit ISR while another strain of<br />

the same species cannot. It is encouraging that implementation of ISR by<br />

endophytic bacilli is beginning, even while some basic questions remain to<br />

be answered.


50 J.W. Kloepper, C.-M. Ryu<br />

<strong>References</strong><br />

Bashan Y, de-Bashan LE (2002) Reduction of bacterial speck (Pseudomonas syringae pv.<br />

tomato) of tomato by combined treatments of plant growth-promoting bacterium,<br />

Azospirillum brasilense, streptomycin sulfate, and chemo-thermal seed treatment. Eur<br />

J Plant Pathol 108:821–829<br />

Benhamou N, Kloepper JW, Quadt-Hallmann A, Tuzun S (1996) Induction of defense-related<br />

ultrastructural modifications in pea root tissues inoculated with endophytic bacteria.<br />

Plant Physiol 112:919–929<br />

Benhamou N, Kloepper JW, Tuzun S (1998) Induction of resistance against Fusarium wilt of<br />

tomato by combination of chitosan with an endophytic bacterial strain: ultrastructure<br />

and cytochemistry of the host response. Planta 204:153–168<br />

Chen C, Bauske EM, Musson G, Rodríguez-Kábana R, Kloepper JW (1995) Biological control<br />

of Fusarium wilt of cotton by use of endophytic bacteria. Biol Control 5:83–91<br />

Enebak SA, Carey WA (2000) Evidence for induced systemic protection to fusiform rust in<br />

loblolly pine by plant growth-promoting rhizobacteria. Plant Dis 84:306–308<br />

Hallman J, Quadt-Hallmann A, Mahaffee WF, Kloepper JW (1997) Bacterial endophytes in<br />

agricultural crops. Can J Microbiol 43:895–914<br />

Iavicoli A, Boutel E, Buchala A, Métraux J-P (2003) Induced systemic resistance in Arabidopsis<br />

thaliana in response to root inoculation with Pseudomonas fluorescens CHA0.<br />

Mol Plant Microb Interact 16:851–858<br />

Jeun YC, Park KS, Kim CH, Fowler WD, Kloepper JW (2004) Cytological observations<br />

of cucumber plants during induced resistance elicited by rhizobacteria. Biol Control<br />

29:34–42<br />

Jetiyanon K, Kloepper JW (2002) Mixtures of plant growth-promoting rhizobacteria for<br />

induction of systemic resistance against multiple plant diseases. Biol Control 24:285–291<br />

Jetiyanon K, Fowler WD, Kloepper JW (2003) Broad-spectrum protection against several<br />

pathogens by PGPR mixtures under field conditions. Plant Dis 87:1390–1394<br />

Kloepper JW, Wei G, Tuzun S (1992) Rhizosphere population dynamics and internal colonization<br />

of cucumber by plant growth-promoting rhizobacteria which induce systemic<br />

resistance to Colletotrichum orbiculare In:TjamosES,PapavizasGC,CookRJ(eds)<br />

Biological control of plant diseases, Plenum, New York, pp 185–191<br />

Kloepper JW, Reddy MS, Rodríguez-Kabana R, Kenney DS, Kokalis-Burelle N, Martinez-<br />

Ochoa N (2004) Application for rhizobacteria in transplant production and yield enhancement.<br />

Acta Hortic 631:219–229<br />

Liu L, Kloepper JW, Tuzun S (1995a) Induction of systemic resistance in cucumber against<br />

Fusarium wilt by plant growth-promoting rhizobacteria. Phytopathology 85:695–698<br />

Liu L, Kloepper JW, Tuzun S (1995b) Induction of systemic resistance in cucumber against<br />

bacterial angular leaf spot by plant growth-promoting rhizobacteria. Phytopathology<br />

85:843–847<br />

Liu L, Kloepper JW, Tuzun S (1995c) Induction of systemic resistance in cucumber by plant<br />

growth-promoting rhizobacteria: duration of protection and effect of host resistance on<br />

protection and root colonization. Phytopathology 85:1064–1068<br />

Maurhofer M, Hase C, Meuwly P, Métraux J-P, Defago G (1994) Induction of systemic<br />

resistance of tobacco to tobacco necrosis virus by the root-colonizing Pseudomonas<br />

fluorescens strain CHA0: influence of the gacA-gene and of pyoverdine production.<br />

Phytopathology 84:139–146<br />

Maurhofer M, Reimmann C, Schmidli-Sachere P, Heeb S, Haas D, Defago G (1998) Salicylic<br />

acid biosynthesis genes expressed in Pseudomonas fluorescens strain P3 improve the induction<br />

of systemic resistance in tobacco against tobacco necrosis virus. Phytopathology<br />

88:678–684


3 Bacterial Endophytes as Elicitors of Induced Systemic Resistance 51<br />

Murphy JF, Zehnder GW, Schuster DJ, Sikora EJ, Polston JE, Kloepper JW (2000) Plant<br />

growth-promoting rhizobacterial mediated protection in tomato against tomato mottle<br />

virus. Plant Dis 84:779–784<br />

Murphy JF, Reddy MS, Ryu C-M, Kloepper JW, Li R (2003) Rhizobacteria-mediated growth<br />

promotion of tomato leads to protection against Cucumber mosaic virus.Phytopathology<br />

93:1301–1307<br />

Park KS, Kloepper JW (2000) Activation of PR-1a promoter by rhizobacteria that induce<br />

systemic resistance in tobacco against Pseudomonas syringae pv. tabaci. BiolControl<br />

18:2–9<br />

Pieterse CMJ, van Wees SCM, van Pelt JA, Knoester M, Laan R, Gerrits H, Weisbeek, PJ, van<br />

Loon LC (1998) A novel signaling pathway controlling induced systemic resistance in<br />

Arabidopsis. Plant Cell 10:1571–1580<br />

Press CM, Wilson M, Tuzun S, Kloepper JW (1997) Salicylic acid produced by Serratia<br />

marcescens 90-166 is not the primary determinant of induced systemic resistance in<br />

cucumber or tobacco. Mol Plant Microb Interact 10:761–768<br />

Press CM, Loper JE, Kloepper JW (2001) Role of iron in rhizobacteria-mediated induced<br />

systemic resistance. Phytopathology 91:593–598<br />

Quadt-Hallmann A, Kloepper JW (1996) Immunological detection and localization of the<br />

cotton endophyte Enterobacter asburiae JM22 in different plant species. Can J Microbiol<br />

42:1144–1154<br />

Quadt-Hallmann A, Hallmann J, Kloepper JW (1997) Bacterial endophytes in cotton: location<br />

and interaction with other plant-associated bacteria. Can J Microbiol 43:254–259<br />

Raupach GS, Liu L, Murphy JF, Tuzun S, Kloepper JW (1996) Induced systemic resistance<br />

in cucumber and tomato against cucumber mosaic cucumovirus using plant growthpromoting<br />

rhizobacteria (PGPR). Plant Dis 80:891–894<br />

Ryu C-M, Farag MA, Hu C-H, Reddy MS, Wei H-X, Paré PW, Kloepper JW (2003a) Bacterial<br />

volatiles promote growth in Arabidopsis. Proc Natl Acad Sci USA 100:4927–4932<br />

Ryu C-M, Hu C-H, Reddy MS, Kloepper JW (2003b) Different signaling pathways of induced<br />

resistance by rhizobacteria in Arabidopsis thaliana against two pathovars of<br />

Pseudomonas syringae. New Phytol 160:413–420<br />

Ryu C-M, Farag MA, Hu C-H, Munagala SR, Kloepper JW, Paré PW (2004a) Bacterial volatiles<br />

induce systemic resistance in Arabidopsis. Plant Physiol 134:1017–1026<br />

Ryu C-M, Murphy JF, Mysore KS, Kloepper JW (2004b) Plant growth-promoting rhizobacteria<br />

systemically protect Arabidopsis thaliana against Cucumber mosaic virus by a salicylic<br />

acid and NPR1-independent and jasmonic acid-dependent signaling pathway.<br />

Plant J 39:381–392<br />

Wei G, Kloepper JW, Tuzun S (1991) Induction of systemic resistance of cucumber to<br />

Colletotrichum orbiculare by select strains of plant growth-promoting rhizobacteria.<br />

Phytopathology 81:1508–1512<br />

Wei G, Kloepper JW, Tuzun S (1996) Induced systemic resistance to cucumber diseases and<br />

increased plant growth by plant growth-promoting rhizobacteria under field conditions.<br />

Phytopathology 86:221–224<br />

Yan Z, Reddy MS, Ryu C-M, McInroy JA, Wilson M, Kloepper JW (2002) Induced systemic<br />

resistance against tomato late blight elicited by plant growth-promoting rhizobacteria.<br />

Phytopathology 92:1329–1333<br />

Zehnder G, Kloepper J, Tuzun S, Yao C, Wei G, Chambliss O, Shelby R (1997a) Insect feeding<br />

on cucumber mediated by rhizobacteria-induced plant resistance. Entomol Exp Appl<br />

83:81–85<br />

Zehnder G, Kloepper J, Yao C, Wei G (1997b) Induction of systemic resistance in cucumber<br />

against cucumber beetles (Coleoptera: Chrysomelidae) by plant growth-promoting<br />

rhizobacteria. J Econ Entomol 90:391–396


52 J.W. Kloepper, C.-M. Ryu<br />

Zehnder GW, Yao C, Murphy JF, Sikora EJ, Kloepper JW (2000) Induction of resistance in<br />

tomato against cucumber mosaic cucumovirus by plant growth-promoting rhizobacteria.<br />

BioControl 45:127–137<br />

Zehnder GW, Murphy JF, Sikora EJ, Kloepper JW (2001) Application of rhizobacteria for<br />

induced resistance. Eur J Plant Pathol 107:39–50<br />

Zhang S, Moyne A-L, Reddy MS, Kloepper JW (2002a) The role of salicylic acid in induced<br />

systemic resistance elicited by plant growth-promoting rhizobacteria against blue mold<br />

of tobacco. Biol Control 25:288–296<br />

Zhang S, Reddy MS, Kloepper JW (2002b) Development of assays for assessing induced systemic<br />

resistance by plant growth-promoting rhizobacteria against blue mold of tobacco.<br />

Biol Control 23:79–86<br />

Zhang S, Reddy MS, Kloepper JW (2004) Tobacco growth enhancement and blue mold<br />

disease protection by rhizobacteria: Relationship between plant growth promotion and<br />

systemic disease protection by PGPR strain 90–166. Plant Soil 262:277–288


4<br />

Control of Plant Pathogenic Fungi<br />

with Bacterial Endophytes<br />

Gabriele Berg, Johannes Hallmann<br />

4.1<br />

Introduction<br />

Interest in biological control has increased over the past years, driven by<br />

the need for alternatives to chemicals – which have often lost their activity<br />

due to the development of resistant pathogen populations – and<br />

to public pressure to develop production systems favourable to the environment<br />

(Whipps 2001). In this respect antagonistic bacteria provide an<br />

environmentally sound alternative to protect plants against attack by fungal<br />

pathogens (Whipps 1997; Bloemberg and Lugtenberg 2001). In the past,<br />

rhizosphere bacteria have been shown to be effective antagonists against<br />

a broad spectrum of fungal pathogens (Weller 1988; Emmert and Handelsman<br />

1999; Kurze et al. 2001). More recent studies have indicated that<br />

bacteria colonising the root interior can even improve plant growth and<br />

plant health (Frommel et al. 1991; Sturz et al. 1999; see Chap. 3 by Kloepper<br />

and Ryu), and seem to be excellent candidates for use as biological control<br />

agents (BCAs) (Chen et al. 1995; Sturz et al. 1997; Downing and Thomson<br />

2000; Sturz et al. 2000; Adhikari et al. 2001; Tjamos et al. 2004; see Chap. 3<br />

by Kloepper and Ryu).<br />

Besides induced resistance (see Chap. 3 by Kloepper and Ryu), little is<br />

known about other mechanisms used by antagonistic endophytic bacteria<br />

towards fungal pathogens, such as antibiosis, competition and lysis. Furthermore,<br />

endophytic bacteria are known to promote plant growth by the<br />

production of plant hormones, enhanced nutrient availability and nitrogen<br />

fixation (Whipps 2001; Hurek and Reinhold-Hurek 2003). For example,<br />

plant hormones produced by endophytic bacteria seem to be essential for<br />

bryophyte development (Hornschuh et al. 2002).<br />

So far, most information about the community structure of endophytic<br />

bacteria with antagonistic properties has been obtained using cultivationdependent<br />

approaches [Chen et al. 1995; Sturz et al. 1999; see Chaps. 2<br />

Gabriele Berg: Graz University of Biotechnology, Department of Environmental Technology,<br />

Petersgasse 12, 8010 Graz, Austria, E-mail: gabriele.berg@TUGraz.at<br />

Johannes Hallmann: Biologische Bundesanstalt für Land- und Forstwirtschaft, Institut für<br />

Nematologie und Wirbeltierkunde, Toppheideweg 88, 48161 Münster, Germany<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


54 G. Berg, J. Hallmann<br />

(Hallmann and Berg) and 15 (Schulz)]. However, recently developed cultivation-independent<br />

methods (reviewed by Smalla 2004) analyse endophytic<br />

communities directly from root tissue. Techniques using bacterial<br />

DNAprovideinformationonthestructuraldiversityoftheentireendophytic<br />

community, while techniques using bacterial RNA identify metabolically<br />

active bacteria, e.g. those which are of interest after pathogen infection.<br />

Primers are available that specifically recognise bacterial taxa with<br />

high percentages of antagonistic strains, such as Pseudomonas, Serratia<br />

or Burkholderia (Widmer et al. 1998; Salles et al. 2001). Unfortunately,<br />

the methodology does not distinguish between antagonistic and nonantagonistic<br />

strains. However, genes involved in bacterial antagonism,<br />

such as those involved in the expression of antibiotics, siderophores or<br />

phytohormones, are now being cloned and in the near future may lead to<br />

the development of primers targeting antagonism and specific microarrays<br />

(Raaijmakers et al. 1997; De Souza and Raaijmakers 2003; Zhou<br />

2003). Overall, a polyphasic approach combining cultivation-dependent<br />

and cultivation-independent methods is recommended to best gain insights<br />

into the dynamics of antagonistic endophytic communities as well<br />

as plant/endophyte/pathogen interactions (Garbeva et al. 2001; Krechel et<br />

al. 2002; Reiter et al. 2002; Sessitsch et al. 2004). Understanding those interactions<br />

provides a platform from which to develop endophytic bacteria<br />

as biocontrol agents.<br />

However, formulation, application, risk assessment, fruit quality and<br />

potential side-effects are further questions that need to be properly answered<br />

before endophytic bacteria can be released as BCAs. This chapter<br />

discusses four key aspects of biological control of fungal pathogens using<br />

endophytic bacteria: (1) the spectrum of indigenous bacterial antagonists<br />

in plant roots, (2) modes of action, (3) use of BCAs, and (4) strategies to<br />

enhance biocontrol efficiency.<br />

4.2<br />

Spectrum of Indigenous Endophytic Bacteria with<br />

Antagonistic Potential Towards Fungal Plant Pathogens<br />

Antagonists are naturally occurring organisms with the potential to interfere<br />

with pathogen infection, growth, and survival (Chernin and Chet<br />

2002). A better understanding of the spectrum of indigenous antagonistic<br />

bacteria will (1) increase our knowledge of plant/endophyte interactions,<br />

(2) facilitate screening efforts for effective biocontrol organisms, (3) allow<br />

breeding of cultivars supporting a high level of antagonistic bacteria, and<br />

(4) may even lead to management strategies for increasing the antagonistic<br />

potential of endophytic bacteria. This chapter focuses on the spectrum of


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 55<br />

antagonistic endophytic bacteria, and discusses the factors that influence<br />

and regulate them.<br />

Antagonistic communities can be studied from a quantitative or qualitative<br />

perspective. Commonly used methods for bacterial identification<br />

include morphological and physiological characterisation, fatty methylester<br />

analysis and molecular techniques based on specific primers and/or<br />

(partial) sequencing of 16S rDNA. While the first method is limited to<br />

culturable bacteria, the latter method can also identify non-culturable endophytic<br />

bacteria, e.g. in a clone library. Antagonistic activity of endophytic<br />

bacteriaisgenerallytestedbyinvitroinhibitionoffungalpathogensindual<br />

cultures and then confirmed in bioassays on host plants. Quantitative analysis<br />

to detect antagonistic bacteria is time-consuming as all the endophytic<br />

bacteria have to be screened for their antagonistic potential. Primers developed<br />

for gene sequences involved in antagonistic activity may, in the<br />

future, be able to recognise antagonists without cultivation.<br />

Using cultivation-dependent methods, the proportion of antagonistic<br />

endophytes can vary between 0%, as shown for the pathosystem Phytophthora<br />

cactorum–potato (Sessitsch et al. 2004) and 50%, for Verticillium<br />

longisporum–oilseed rape (Graner et al. 2003) (Table 4.1). Major factors<br />

determining the percentage of antagonists are most likely plant species,<br />

pathogen infestation, habitat and vegetation period (Sessitsch et al. 2004;<br />

Berg et al. 2005). In conclusion, these data confirm that a significant portion<br />

of the indigenous endophytic bacteria in plant roots have antagonistic<br />

potential towards fungal pathogens.<br />

Table 4.1. Proportion of antagonistic endophytic bacteria in different pathosystems<br />

Plant Plant pathogens Proportion of<br />

antagonists (%)<br />

Reference<br />

Rice Achlya klebsiana 25 Adhikari et al. (2001)<br />

Pythium spinosum 75<br />

Potato Verticillium dahliae<br />

Rhizoctonia solani<br />

9 Krechel et al. (2002)<br />

Potato Verticillium dahliae 13 Berg et al. (2004)<br />

Rhizoctonia solani 9.7<br />

Potato Verticillium dahliae 2 Sessitsch et al. (2004)<br />

Rhizoctonia solani 3<br />

Phytophthora cactorum 0<br />

Streptomyces scabies 43<br />

Xanthomonas campestris 29<br />

Oilseed rape Verticillium longisporum 50 Graner et al. (2003)<br />

Tomato Verticillium dahliae 12 Tjamos et al. (2004)


56 G. Berg, J. Hallmann<br />

But what are the main bacterial species conferring antagonism towards<br />

fungal plant pathogens? Although a high diversity of antagonistic endophytic<br />

bacteria is found in general, some genera harbour more antagonistic<br />

strains than others, e.g. Bacillus, Curtobacterium, Methylobacterium,<br />

Paenibacillus, Pseudomonas, Serratia, Stenotrophomonas and Streptomyces<br />

(Sturz et al. 1999; Garbeva et al. 2001; Krechel et al. 2002; Reiter et al. 2002;<br />

Sessitsch et al. 2004). Antagonistic species have been isolated from a number<br />

of different plant species, but most studies have been done on potato.<br />

Table 4.2 lists antagonistic species found in potato in five studies. Surprisingly,<br />

a total of 51 different species comprising 27 genera were identified!<br />

The proportion and composition of indigenous endophytic bacteria with<br />

antagonistic capacity is influenced by a variety of biotic and abiotic factors,<br />

with the plant itself being a major factor (Germida et al. 1998). As shown by<br />

Table 4.2. Endophytic bacterial species of potato with antagonistic properties towards plant<br />

pathogenic fungi a<br />

Species with antagonistic properties<br />

Agrobacterium tumefaciens<br />

Amycolatopsis meditarranei<br />

Arthrobacter ilicis<br />

Bacillus aquamarinus<br />

Bacillus cereus<br />

Bacillus megaterium<br />

Clavibacter michigenensis<br />

Curtobacterium albidum<br />

Curtobacterium flaccumfaciens<br />

Curtobacterium luteum<br />

Erwinia persicinus<br />

Flavobacterium sp.<br />

Frateuria aurantia<br />

Frigoribacterium sp.<br />

Kingella kingae<br />

Kitasatosporia cystargenia<br />

Methylobacterium sp.<br />

Micrococcus varians<br />

Paenibacillus sp.<br />

Pantoaea agglomerans<br />

Pantoaea anantis<br />

Pseudomonas cichorii<br />

Pseudomonas corrugata<br />

Pseudomonas fluorescens<br />

Pseudomonas graminis<br />

Pseudomonas migulae<br />

Pseudomonas orientalis<br />

Pseudomonas putida<br />

Pseudomonas rhodesiae<br />

Pseudomonas reactans<br />

Pseudomonas straminea<br />

Pseudomonas synthaxa<br />

Pseudomonas syringae<br />

Pseudomonas tolaasii<br />

Pseudomonas veronii<br />

Psychrobacter immobilis<br />

Ralstonia paucula<br />

Rhizobium meliloti<br />

Rhizomonas suberifaciens<br />

Sphigobacterium thalophilum<br />

Sphingomonas adhaesiva<br />

Stenotrophomonas maltophilia<br />

Streptomyces turgidiscabies<br />

Streptomyces bottropensis<br />

Streptomyces diastatochromogenes<br />

Streptomyces galilaeus<br />

Streptomyces griseus<br />

Streptomyces lavendulae<br />

Streptomyces setonii<br />

Streptomyces turgidiscabies<br />

Xanthomonas campestris<br />

Xanthomonas oryzae<br />

a According to Sturz et al. (1999), Krechel<br />

et al. (2002), Reiter et al. (2002), Sessitsch<br />

et al. (2004), and A. Krechel et al., unpublished<br />

data


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 57<br />

Zinniel et al. (2002) for endophytic bacteria in above-ground plant organs,<br />

the bacterial spectrum of 4 agronomic crop species and 27 prairie plant<br />

species varied greatly. A similar effect of the plant species on the bacterial<br />

spectrum can also be expected for antagonistic root endophytes. There is<br />

even an effect of the cultivar or plant genotype on the bacterial spectrum<br />

(see Chap. 2 by Hallmann and Berg). For example, the oilseed rape cultivar<br />

‘Express’, which is tolerant to V. longisporum,containedahigherproportion<br />

of bacteria with proteolytic, cellulolytic and phosphatase activity than<br />

the susceptible cultivar ‘Libraska’ (Graner et al. 2003). Modern wheat cultivars<br />

have a more diverse endophytic community than ancient land races,<br />

and are more aggressively colonised by endophytic pseudomonads that<br />

produce antifungal metabolites (Germida and Siciliano 2001). Breeder selection<br />

for higher productivity might have selected plants that support an<br />

endophytic microflora antagonistic to fungal pathogens. Finally, the plant<br />

tissue itself can also vary in its antagonistic bacteria. Sturz et al. (1999)<br />

made the observation that the outer tissue of a potato tuber yielded higher<br />

relative densities of endophytic bacteria antagonistic to soilborne fungal<br />

pathogens than deeper layers of the potato. Furthermore, the antifungal<br />

potential of bacterial endophytes was highest for isolates recovered from<br />

the outermost layer of the tuber. The authors assumed that, in certain<br />

communities of endophytic bacteria, antagonism against fungal pathogens<br />

may be related to bacterial adaptation to location within a host plant, and<br />

may be tissue-type and tissue-site specific, indicating that in plant tissue<br />

exposed to the soil, antagonistic endophytes become more prominent. Is<br />

this a result of coevolution or a process controlled by the plant itself? Such<br />

questions still await proper answers.<br />

Besides the plant itself, several biotic factors affect endophytic communities<br />

and the proportion of antagonists (Reiter et al. 2002; Sessitsch et al.<br />

2002, 2004). In this context, it was shown that pathogen stress had a greater<br />

impact than plant genotype on bacterial diversity. Furthermore, Sessitsch<br />

et al. (2004) found a different spectrum of antagonistic bacteria in good<br />

and poor growing potatoes in the field.<br />

Fluctuations in antagonistic bacterial communities of plant roots may<br />

also be caused by abiotic factors such as temperature, rainfall, cropping<br />

practice or soil amendments (Mocali et al. 2003). Plants are able to select<br />

specific bacterial genotypes in response to soil conditions (Siciliano et<br />

al. 2001). Therefore, the bacterial community in the soil is an important<br />

factor affecting the composition of indigenous endophytic bacteria. Soil<br />

amendments can modify the bacterial spectrum in the soil as well as in the<br />

roots (Hallmann et al. 1999) and enhance the efficacy of biocontrol agents<br />

(Ahmed et al. 2003).<br />

In conclusion, the spectrum and diversity of endophytic bacterial communities<br />

in plant roots is never a stable scenario, rather it has its own


58 G. Berg, J. Hallmann<br />

dynamics in response to biotic as well as abiotic factors. This offers new<br />

opportunities for managing the indigenous spectrum of antagonistic endophytic<br />

bacteria to increase the benefits to the plant by means of plant<br />

breeding, soil amendments or application of endophytic BCAs.<br />

4.3<br />

Mode of Action of Antagonistic Bacteria<br />

Modes of action of antagonistic towards fungal pathogens have been intensively<br />

studied for plant growth-promoting rhizobacteria, as reviewed<br />

by Fravel (1988), Whipps (2001), Lugtenberg et al. (2001), and Bloemberg<br />

et al. (2001). Presumably, endophytic bacteria use similar mechanisms for<br />

the control of fungal plant pathogens. However, their hidden life within<br />

the plant tissue makes it much more difficult to study such mechanisms<br />

(see Chap. 18 by Bloemberg and Camacho Carvajal). Furthermore, it is<br />

often difficult to distinguish between direct antagonism (such as antibiosis),<br />

competition and lysis, and indirect mechanisms such as induced resistance<br />

and improved plant growth (see also Chap. 3 by Kloepper and<br />

Ryu).<br />

4.3.1<br />

Antibiosis<br />

Antibiosis describes the ability of an endophytic bacterium to inhibit<br />

pathogen growth by the production of antibiotics or toxins. Although<br />

the vast majority of endophytic bacteria show antibiosis against fungal<br />

pathogens in vitro (Krechel et al. 2002; Sturz et al. 1999), very little is known<br />

about the significance of antibiosis controlling fungal pathogens within the<br />

root tissue. Examples for antifungal substances released by endophytic<br />

bacteria include iturin A (produced by Bacillus subtilis) and pyrrolnitrin<br />

(produced by Serratia plymuthica) (Cho et al. 2002). Further support for<br />

the hypothesis that these antifungal metabolites represent the underlying<br />

mechanisms in situ could be achieved by antibiotic-deficient mutants that<br />

fail to express biocontrol activity. Unfortunately, no such studies have yet<br />

been carried out.<br />

However, just as microbial antagonists utilise a diverse arsenal of mechanisms<br />

to dominate interactions with fungal pathogens, pathogens have<br />

surprisingly diverse responses to counteract these antagonisms (Duffy and<br />

Défago 1997). These responses include detoxification, antibiotic resistance,<br />

active efflux of antibiotics, and repression of biosynthetic genes expressing<br />

proteins involved in biocontrol (Duffy et al. 2003). Again, most work in


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 59<br />

this area has been carried out on rhizosphere bacteria, and almost nothing<br />

is known about the regulation of antifungal metabolites expressed by<br />

endophytic bacteria. However, since antibiosis seems to be a mechanism<br />

of fungal control used by endophytic bacteria, metabolites released by the<br />

bacteria into plant tissue must be carefully monitored to ensure that they<br />

pose no risk regarding fruit quality and consumer health.<br />

4.3.2<br />

Competition<br />

Competition is considered an important factor in the control of fungal<br />

pathogens by endophytic bacteria, since both organisms colonise similar<br />

niches and utilise the same nutrients. Conclusive data demonstrating<br />

competition as a major control mechanism of endophytic bacteria are still<br />

lacking. Work on rhizosphere bacteria has shown that, under iron-limiting<br />

conditions, bacteria produce siderophores with high affinity for ferric iron.<br />

By binding available iron these bacteria deprive fungal pathogens of iron,<br />

thus restricting their growth (O’Sullivan and O’Gara 1992). Siderophores<br />

are also commonly produced by endophytic bacteria (Krechel et al. 2002),<br />

indicating that similar mechanisms may occur in the endorhiza.<br />

4.3.3<br />

Lysis<br />

Cell wall lysis is another potential mechanism whereby endophytic bacteria<br />

can control fungal pathogens. This mechanism is well established in the<br />

biocontrol of fungal pathogens by rhizosphere bacteria. Endophytic bacteria<br />

isolated from potato roots express high levels of hydrolytic enzymes<br />

such as cellulase, chitinase and glucanase (Krechel et al. 2002). Pleban et al.<br />

(1997) analysed the importance of lytic enzymes in antagonism of Bacillus<br />

cereus strain 65 towards the soilborne fungal pathogen Rhizoctonia<br />

solani. B. cereus strain 65, originally isolated from surface-sterilised seeds<br />

of Sinapis arvensis, was shown to excrete a chitinase of 36 kDa, responsible<br />

for the observed protection of cotton seedlings from root rot disease caused<br />

by R. solani. Additionally, chitinolytic Bacillus subtilis strains were able to<br />

reduce symptoms of Verticillium dahliae in several host plants (Tjamos et<br />

al. 2004). An endophytic chitinase-producing isolate of Actinoplanes missouriensis<br />

and its culture filtrates were shown to suppress Plectosporium<br />

tabacinum on lupins (El-Tarabily 2003). The importance of hydrolytic enzymes<br />

other than chitinases as biocontrol mechanisms is still unknown.


60 G. Berg, J. Hallmann<br />

4.3.4<br />

Induction of Plant Defence Mechanisms<br />

Induction of plant defence mechanisms by endophytic bacteria plays a major<br />

role in suppression of fungal plant pathogens and therefore is covered<br />

in a separate chapter (Chap. 3 by Kloepper and Ryu).<br />

4.3.5<br />

Plant Growth<br />

Plant growth is a factor that is indirectly involved in pathogen defence.<br />

Plants with vigorous growth, such as cucumber, can sometimes outgrow<br />

disease by fungal pathogens such as powdery mildew. Therefore, plant<br />

growth promotion by endophytic bacteria indirectly affects the pathogenicity<br />

of fungal pathogens. Nejad and Johnson (2000) described isolates of endophytic<br />

bacteria that significantly improved seed germination and plant<br />

growth of oilseed rape and tomato. Plant growth promotionmediated by endophytic<br />

bacteria may be exerted by several mechanisms, e.g. production of<br />

plant growth hormones, synthesis of siderophores, nitrogen fixation, solubilisation<br />

of minerals such as phosphorous, or via enzymatic activities, for<br />

example suppression of ethylene by 1-aminocyclopropane-1-carboxylate<br />

(ACC) deaminase (Chernin and Chet 2002). Strains of Pseudomonas, Enterobacter,<br />

Staphylococcus, Azotobacter and Azospirillum produce plant<br />

growth regulators such as ethylene, auxins or cytokinins, which are assumed<br />

to promote plant growth (Arshad and Frankenberger 1991; Leifert<br />

et al. 1994). However, in the past, most interest has focussed on the fixation<br />

of atmospheric nitrogen by free-living endophytic bacteria, especially of<br />

diazotrophs (Döbereiner and Pedrosa 1987; Hecht-Buchholz 1998; Estrada<br />

et al. 2002; Hurek and Reinhold-Hurek 2003).<br />

Overall, mechanisms of fungal control by endophytic bacteria may<br />

act synergistically, and individual endophytes quite often exhibit several<br />

modes of action. However, the expression of antifungal mechanisms is<br />

strain-specific (Neiendam-Nielson et al. 1998; Berg 2000; Berg et al. 2002)<br />

and most likely under the control of several biotic as well as abiotic factors.<br />

A better understanding of the underlying mechanisms has significant<br />

relevance for the optimisation of biocontrol strategies (see Sect. 4.5).<br />

4.4<br />

Control Potential of Endophytic Bacteria<br />

A broad spectrum of endophytic bacteria has been described to control<br />

fungal plant pathogens on different plant species (Table 4.3). The major-


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 61<br />

ity of antagonistic endophytic bacteria are Gram-negative and belong to<br />

the group of fluorescent pseudomonads, which are effective BCAs (Bloemberg<br />

and Lugtenberg 2001; Whipps 2001). Fluorescent pseudomonads are<br />

common members of the endorhiza, making them ideal candidates for biological<br />

control measures. Antagonistic isolates have been reported to occur<br />

in Pseudomonas chlororaphis, Pseudomonas fluorescens, Pseudomonas<br />

graminis, Pseudomonas putida, Pseudomonas tolaasii and Pseudomonas<br />

veronii (Table 4.3). Control potential in terms of reduced disease severity<br />

can approach 80% under greenhouse conditions (Pleban et al. 1995). Besides<br />

pseudomonads, other Gram-negative species with biocontrol activity<br />

against fungal pathogens are Phyllobacterium rubiacearum, Burkholderia<br />

solanacearum (Chen et al. 1995), Sphingomonas trueperi (Adhikari et al.<br />

2001) and Serratia plymuthica (Faltin et al. 2004).<br />

Gram-positive bacterial isolates with antagonistic properties are commonly<br />

found within the genus Bacillus (Emmert and Handelsman 1999).<br />

Bacillus spp. occur mainly in the soil/rhizosphere but have also been reported<br />

as endophytes of several plant species (see Chap. 2 by Hallmann<br />

and Berg). As summarized in Table 4.3, endophytic isolates of Bacillus,and<br />

the closely related genus Paenibacillus, havebeenshowntosignificantly<br />

control many fungal diseases. Gram-positive biocontrol bacteria other than<br />

bacilli include actinomycetes such as Streptomyces, Microbispora or Nocardioides<br />

(Coombs et al. 2004). For example, the streptomycete Actinoplanes<br />

missouriensis gave excellent control of Plectosporium tabacinum, the causal<br />

agent of lupin root rot (El-Tarabily 2003).<br />

Most studies have been carried out under greenhouse conditions. However,<br />

a few studies have achieved similar biocontrol effects under field<br />

conditions (Tjamos et al. 2004; Faltin et al. 2004). For example, the endophytic<br />

isolate Pseudomonas trivialis 3Re2-7 significantly reduced disease<br />

incidence of Rhizoctonia solani onlettuceandsugarbeetby40%and86%,<br />

respectively (Faltin et al. 2004).<br />

These examples illustrate the high potential of endophytic bacteria in<br />

fungal pathogen control. However, further fieldwork is required to confirm<br />

their control efficacy in different climatic regions and under different<br />

growth conditions. Formulations and applications that meet common<br />

farming practice still need to be developed. For promising BCAs, strategies<br />

that enhance overall control efficacy should be explored.<br />

4.5<br />

Enhancing Biocontrol Efficiency<br />

It is a well-accepted observation that biological control under field conditions<br />

is often inconsistent (Weller 1988). Therefore, enhancing the efficacy


62 G. Berg, J. Hallmann<br />

Table 4.3. Examples of antagonistic<br />

Pathosystem BCA Trial Application Results Reference<br />

Pleban et al.<br />

1995, 1997<br />

Cotton, Bean<br />

(Rhizoctonia solani)<br />

Pleban et al.<br />

1995<br />

Chen et al.<br />

1995<br />

Bean<br />

(Sclerotium rolfsii)<br />

Cotton<br />

(Fusarium oxysporum<br />

f. sp. vasinfectum)<br />

Sharma and<br />

Nowak 1998<br />

Downing and<br />

Thomson<br />

2000<br />

Bacillus cereus<br />

Greenhouse Root incubation in Reduction of disease incidence<br />

Bacillus subtilis<br />

bacterial suspension approx. 50%<br />

Bacillus pumilus<br />

Bacillus subtilis<br />

Greenhouse Root incubation in Reduction of disease incidence<br />

Bacillus cereus<br />

bacterial suspension of 70–80%<br />

Aureobacterium saperdae Pot experiment Pierced with a needle Reduction of disease severity<br />

Bacillus pumilus<br />

and plant growth promotion<br />

Phyllobacterium rubiacearum<br />

Pseudomonas putida<br />

Burkholderia solanacearum<br />

Pseudomonas Growth Bacterization of tissue Reduction of disease severity<br />

chamber plantlets or seedlings and plant growth promotion<br />

Pseudomonas fluorescens Plant growth Soil drenching<br />

Rif1::tc4 (with chiA gene) chamber 107 cfu ml−1 Reduction of disease incidence<br />

up to 48%, only if introduced as<br />

Introduction into the endophytes<br />

Tomato<br />

(Verticillium dahliae)<br />

Bean<br />

(Rhizoctonia solani)<br />

plant<br />

Adhikari et al.<br />

2001<br />

Reduction of disease incidence<br />

by 50–73%; plant growth<br />

promotion (plant height and<br />

Pot experiments Rice seeds were<br />

soakedfor1hin<br />

a bacterial suspension<br />

108 cfu ml−1 Pseudomonas fluorescens<br />

Pseudomonas tolaasii<br />

Pseudomonas veronii<br />

Rice<br />

(Achlya klebsiana,<br />

Pythium spinosum)<br />

dry weight)<br />

Sphingomonas trueperi<br />

Krechel et al.<br />

2002<br />

Seed treatment Reduction of fungal growth in<br />

vitro, of nematode infestation<br />

ad planta; plant growth<br />

promotion up to 78%<br />

Growth<br />

chamber<br />

Streptomyces turgidiscabies<br />

Kitasatosporia cystargenia<br />

Streptomyces galilaeus<br />

Streptomyces griseus<br />

Pseudomonas graminis<br />

Potato<br />

[Rhizoctonia solani,<br />

Verticillium dahliae,<br />

Meloidogyne incognita<br />

(nematode)]


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 63<br />

Table 4.3. (continued)<br />

Pathosystem BCA Trial Application Results Reference<br />

Bacillus sp. Pot experiments Soil inoculation Reduction of disease incidence Cho et al. 2003<br />

Ahmed et al.<br />

2003<br />

Reduction of disease incidence<br />

of R. solani up to 55% and of<br />

Greenhouse Seed treatments<br />

Root drenching<br />

Bacillus subtilis<br />

Bacillus licheniformis<br />

Balloon flower<br />

(Rhizoctonia solani)<br />

Pepper<br />

(Rhizoctonia solani,<br />

P. capsici up to 55%<br />

Phytophthora capsici)<br />

Tjamos et al.<br />

2004<br />

Reduction of disease severity by<br />

40–70%; yield increase 25%<br />

Root dipping<br />

Soil drenching<br />

107 cfu ml−1 talc-gum<br />

xanthan formulation<br />

Greenhouse<br />

Field<br />

Bacillus spp.<br />

Paenibacillus alvei<br />

Bacillus amiloliquefaciens<br />

Eggplant, potato<br />

(Verticillium dahliae)<br />

Faltin et al.<br />

2004<br />

Reduction of disease severity up<br />

to 76%<br />

Soil drenching<br />

Seed bacterization<br />

108 cfu ml−1 Greenhouse<br />

Field<br />

Serratia plymuthica<br />

Pseudomonas reactans<br />

Potato<br />

(Rhizoctonia solani)<br />

Coombs et al.<br />

2004<br />

Reduction of black lesions<br />

up to 71%<br />

Greenhouse Seed treatments 109 to 10 10 cfu ml−1 Streptomyces spp.<br />

Microbispora spp.<br />

Nocardioides<br />

Wheat<br />

(Gaeumannomyces<br />

graminis var.tritici)


64 G. Berg, J. Hallmann<br />

and consistency of control by endophytic bacteria is a major factor in<br />

determining their future success as BCAs. Current strategies to enhance<br />

biocontrol efficiency include (1) optimised formulation and application<br />

technologies, (2) integrated biocontrol strategies, (3) management of the<br />

indigenous endophytic microflora, and (4) genetic engineering.<br />

The formulation of BCAs has undergone major progress in recent years<br />

(Burghes 1998). Most work has been done on rhizosphere bacteria and<br />

the acquired know-how now needs to be transferred to endophytic bacteria.<br />

Basically, two types of formulation seem to be preferable for endophytic<br />

bacteria: dry products and liquid suspensions. For Gram-positive<br />

endophytes, dry formulations with a long shelf life are already standard<br />

(Emmert and Handelsmann 1999; Whipps 1997) whereas stable formulations<br />

for Gram-negative bacteria are more difficult, due mainly to the lack<br />

of resting spores that can withstand the formulation process. Optimum<br />

formulations should ensure bacterial survival, but also promote bacterial<br />

activity after application. Microcapsules that protect the incorporated bacteria,<br />

while also providing nutrient sources, are an interesting alternative<br />

(Burghes 1998).<br />

The delivery of BCAs into the host plant is a prerequisite for successful<br />

biocontrol. Musson et al. (1995) studied eight methods for delivering<br />

endophytic bacteria into cotton stems and roots: stab-inoculation of bacteria<br />

into stems, soaking seed in bacterial suspensions, soil drench, methyl<br />

cellulose seed coating, foliar spray, bacteria-impregnated granules applied<br />

in-furrow, vacuum infiltration, and the pruned-root dip. Whereas each of<br />

the methods effectively established most of the endophytic bacteria based<br />

on re-isolation studies, none of the methods successfully delivered all 15<br />

strains tested, indicating that the optimum method is strain-specific. For<br />

practical reasons, soil drench, root dipping, and seed coating are the most<br />

promising methods. Standard technology can be used and the bacterium<br />

is delivered directly into the root system where it can immediately colonise<br />

the root and protect the plant against invasion by fungal pathogens.<br />

Fahey et al. (1991) described a seed inoculation technique in which<br />

pressure is applied to infuse the bacterial suspension into imbibed seeds,<br />

followed by redrying of the seeds. The inoculated seeds met the requested<br />

shelf-life requirements of several months, and application of commonly<br />

used fungicides had no adverse impact on bacterial survival or efficacy of<br />

the bacterial inoculation. This leads to another interesting option for biocontrol<br />

enhancement: the combination of endophytic bacteria with other<br />

BCAs or even chemicals. Integrated biological control strategies against<br />

fungal pathogens using a combination of antagonistic endophytes with<br />

complementary modes-of-action and/or colonisation sites, or of endophytes<br />

with synthetic control agents take advantage of synergies and should<br />

be exploited further.


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 65<br />

The naturally occurring bacterial antagonistic potential within a plant is<br />

influenced by several biotic and abiotic factors (see Chap. 2 by Hallmann<br />

and Berg). By changing these factors, the antagonistic potential can be<br />

managed. For example, chitin application not only increased total numbers<br />

of antagonistic Burkholderia cepacia (Hallmann et al. 1999) but, as shown<br />

by Ahmed et al. (2003), also improved control of root rot disease in pepper<br />

by endophytic Bacillus subtilis and B. licheniformis. Furthermore, the chitin<br />

derivate chitosan was successfully used in combination with the endophytic<br />

bacterium Bacillus pumilus to enhance plant resistance towards Fusarium<br />

wilt of tomato (Benhamou et al. 1998). It is hypothesised that chitin as well<br />

as chitosan stimulate the establishment of a chitinolytic microflora, which<br />

then decomposes fungal cell wall chitin.<br />

Biocontrol efficiency might also be improved by breeding plant genotypes<br />

to support a high level of antagonistic endophytes. For example,<br />

plants expressing high levels of N-acylhomoserine lactones, or which are<br />

capable of degrading this important signal molecule of bacterial communication,<br />

have been shown to also influence plant-associated bacteria (Fray<br />

2002). On the other hand, endophytic bacteria might also be promoted<br />

by transgenic means. Promising biocontrol targets for genetic engineering<br />

are gene-regulated factors of endophytic bacteria involved in modes-ofaction,<br />

colonisation potential, survival, fitness, and adaptation to environmental<br />

conditions. For example, the endophytic bacterium Pseudomonas<br />

fluorescens, originally isolated from micropropagated apple plantlets, was<br />

genetically modified by Downing and Thomson (2000) to harbour a gene<br />

encoding the major chitinase of Serratia marcescens. ThegenechiA was<br />

cloned under the control of the tac promoter in the broad-host-range plasmid<br />

pKT240 and the integration vector pJFF350. P. fluorescens carrying<br />

tac-chiA either on the plasmid or integrated into the chromosome significantly<br />

controlled Rhizoctonia solani on beans. Although a promising<br />

approach, genetically modified BCAs still face many restrictions, making<br />

broad-scale application in the near future improbable and difficult.<br />

4.6<br />

Conclusions<br />

Most plants are colonised by a broad spectrum of endophytic bacteria that<br />

are potentially antagonistic towards fungal plant pathogens. This enormous<br />

potential needs to be further explored for its use in modern plant disease<br />

control strategies. This requires not only a better understanding of the<br />

underlying mechanisms and their regulation in response to environmental<br />

factors, but also a more comprehensive picture of what triggers endophytic<br />

colonisation as well as of the population dynamics of antagonistic bacterial


66 G. Berg, J. Hallmann<br />

endophytes within the plant. Continuing research in this area will hopefully<br />

lead to new and innovative concepts for biological control of fungal<br />

pathogens.<br />

<strong>References</strong><br />

Adhikari TB, Joseph CM, Yang G, Phillips DA, Nelson LM (2001) Evaluation of bacteria<br />

isolated from rice for plant growth promotion and biological control of seedling disease<br />

of rice. Can J Microbiol 47:916–924<br />

Ahmed AS, Ezziyyani M, Pérez Sánchez C, Candela ME (2003) Effect of chitin on biological<br />

control activity of Bacillus spp. and Trichoderma harzianum against root rot disease in<br />

pepper (Capsicum annuum) plants. Eur J Plant Pathol 109:633–637<br />

Arshad M, Frankenberger WT (1991) Microbial production of plant hormones. In: Keister<br />

DL, Cregan PB (eds) The rhizosphere and plant growth. Kluwer, Dordrecht,<br />

pp 327–334<br />

Benhamou N, Kloepper JW, Tuzun S (1998) Induction of resistance against Fusarium<br />

wilt of tomato by combination of chitosan with an endophytic bacterial strain. Planta<br />

204:153–168<br />

Berg G (2000) Diversity of antifungal and plant-associated Serratia strains. J Appl Microbiol<br />

88:952–960<br />

Berg G, Roskot N, Steidle A, Eberl L, Zock A, Smalla K (2002) Plant-dependent genotypic and<br />

phenotypic diversity of antagonistic rhizobacteria isolated from different Verticillium<br />

host plants. Appl Environ Microbiol 68:3328–3338<br />

Berg G, Krechel A, Ditz M, Faupel A, Sikora RA, Ulrich A, Hallmann J (2005) Endophytic and<br />

ectophytic potato-associated bacterial communities differ in structure and antagonistic<br />

function against plant pathogenic fungi. FEMS Microbiol Ecol 51:215–229<br />

Bloemberg GV, Lugtenberg BJJ (2001) Molecular basis of plant growth promotion and<br />

biocontrol by rhizobacteria. Curr Opin Plant Biol 4:343–350<br />

Burghes HD (1998) Formulation of biopesticides. Kluwer, Dordrecht<br />

Chen C, Bauske EM, Musson G, Rodríguez-Kábana R, Kloepper JW (1995) Biological control<br />

of Fusarium wilt on cotton by use of endophytic bacteria. Biol Control 5:83–91<br />

Chernin, L, Chet I (2002) Microbial enzymes in biocontrol of plant pathogens and pests. In:<br />

Burns R, Dick R (eds) Enzymes in the environment: activity, ecology, and applications.<br />

Dekker, New York, pp 171–225<br />

Cho SJ, Lim WJ, Hong SY, Park SR, Yun HD (2002) Endophytic colonization of balloon<br />

flower by antifungal strain Bacillus sp. CY22. Biosci Biotech Biochem 67:2132–2138<br />

Coombs JT, Michelson PP, Franco CMM (2004) Evaluation of endophytic actinobacteria as<br />

antagonists of Gaeumannomyces graminis var. tritici in wheat. Biol Control 29:359–366<br />

De Souza JT, Raaijmakers JM (2003) Polymorphisms within the prnD and pltC genes from<br />

pyrrolnitrin and pyoluteorin-producing Pseudomonas and Burkholderia spp. FEMS<br />

Microbiol Ecol 43:21–34<br />

Döbereiner J, Pedrosa FO (1987) Nitrogen-fixing bacteria in non-leguminous crop plants.<br />

Science Tech, Madison, WI<br />

Downing KJ, Thomson JA (2000) Introduction of the Serratia marcescens chiAgeneintoan<br />

endophytic Pseudomonas fluorescensfluorescens for the biocontrol of phytopathogenic<br />

fungi. Can J Microbiol 46:363–369<br />

Duffy BK, Défago G (1997) Fusarium pathogenicity factor blocks antibiotic biosynthesis by<br />

Pseudomonas biocontrol strains. IOBC WPRS Bull 21:145–148


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 67<br />

Duffy B, Schouten A, Raaijmakers JM (2003) Pathogen self-defense: mechanisms to counteract<br />

microbial antagonism. Annu Rev Phytopathol 41:501–538<br />

El-Tarabily KA (2003) An endophytic chitinase-producing isolate of Actinoplanes missouriences,<br />

with potential for biological control of root rot of lupin caused by Plectosporium<br />

tabacinum. Aust J Bot 51:257–266<br />

Emmert EAB, Handelsman J (1999) Biocontrol of plant disease: a (Gram + )positiveperspective.<br />

FEMS Microbiol Lett 171:1–9<br />

Estrada P, Mavingui P, Cournoyer B, Fontaine F, Balandreau J, Caballero-Mellado J (2002)<br />

AN2-fixing endophytic Burkholderia sp. associated with maize plants cultivated in<br />

Mexico. Can J Microbiol 48:285–294<br />

Fahey JW, Dimock MB, Tomasino SF, Taylor JM, Carlson, PS (1991) Genetically engineered<br />

endophytes as biocontrol agents: a case study from industry. In: Andrews JH,<br />

Hirano SS (eds) Microbial ecology of leaves. Springer, Berlin Heidelberg New York,<br />

pp 401–411<br />

Faltin F, Lottmann J, Grosch R, Berg G (2004) Strategy to select and assess antagonistic<br />

bacteria for biological control of Rhizoctonia solani Kühn. Can J Microbiol 50:811–820<br />

Fray RG (2002) Altering plant-microbe interaction through artificially manipulating bacterial<br />

quorum sensing. Ann Bot 89:245–253<br />

Fravel DR (1988) Role of antibiosis in the biocontrol of plant diseases. Annu Rev Phytopathol<br />

26:75–91<br />

Frommel MI, Nowak J, Lazarovits J(1991) Growth enhancement and developmental modificationsofinvitrogrownpotato(Solanum<br />

tuberosum ssp. tuberosum) asaffectedby<br />

anonfluorescentPseudomonas sp. Plant Physiol 96:928–936<br />

Garbeva P, van Overbeek LS, van Vuurde JWL, van Elsas JD (2001) Analysis of endophytic<br />

bacterial communities of potato by plating and denaturing gradient gel electrophoresis<br />

(DGGE) of 16S rDNA based PCR fragments. Microb Ecol 413:69–383<br />

Germida JJ, Siciliano SD (2001) Taxonomic diversity of bacteria associated with the roots<br />

of modern, recent and ancient wheat cultivars. Biol Fertil Soils 33:410–415<br />

Germida JJ, Siciliano SD, Freitas JR de, Seib AM (1998) Diversity of root-associated bacteria<br />

associated with field-grown canola (Brassica napus L.) and wheat (Triticum aestivum L.).<br />

FEMS Microbiol Ecol 26:43–50<br />

Graner G, Persson P, Meijer J, Alstrom S (2003) A study on microbial diversity in different<br />

cultivars of Brassica napus in relation to its wilt pathogen, Verticillium longisporum.<br />

FEMS Microbiol Lett 29:269–276<br />

Hallmann J, Rodríguez-Kábana R, Kloepper JW (1999) Chitin-mediated changes in bacterial<br />

communities of the soil, rhizosphere and within roots of cotton in relation to nematode<br />

control. Soil Biol Biochem 31:551–560<br />

Hecht-Buchholz C (1998) The apoplast-habitat of endophytic nitrogen-fixing bacteria and<br />

their significance for the nitrogen nutrition on nonleguminous plants. Z Pflanzenernähr<br />

Bodenkd 161:509–520<br />

Hornschuh M, Grotha R, Kutschera U (2002) Epiphytic bacteria associated with the<br />

bryophyte Funaria hygrometrica: effectsofMethylobacterium strains on protonema<br />

development. Plant Biol 4:682–687<br />

Hurek T, Reinhold-Hurek B (2003) Azoarcus sp. strain BH72 as a model for nitrogen-fixing<br />

grass endophytes. J Biotechnol 106:169–178<br />

Krechel A, Faupel A, Hallmann J, Ulrich A, Berg G (2002) Potato-associated bacteria and their<br />

antagonistic potential towards plant-pathogenic fungi and the plant-parasitic nematode<br />

Meloidogyne incognita (Kofoid & White) Chitwood. Can J Microbiol 48:772–786<br />

Kurze S, Dahl R, Bahl H, Berg G (2001) Biological control of fungal strawberry diseases by<br />

Serratia HRO-C48. Plant Dis 85:529–534


68 G. Berg, J. Hallmann<br />

Leifert C, Morris CE, Waites WM (1994) Ecology of microbial saprophytes and pathogens<br />

in tissue culture and field-grown plants: reasons for contamination problems in vitro.<br />

Crit Rev Plant Sci 13:139–183<br />

Lugtenberg BJJ, Dekkers L, Bloemberg GV (2001) Molecular determinants of rhizosphere<br />

colonization by Pseudomonas. Annu Rev Phytopathol 39:461–490<br />

Mocali S, Bertelli E, Di Cello F, Mengoni A, Sfalanga A, Viliani F, Caciotti A, Tegli S, Surico G,<br />

Fani R (2003) Fluctuation of bacteria isolated from elm tissues during different seasons<br />

and from different plant organs. Res Microbiol 154:105–114<br />

Musson G, McInroy JA, Kloepper JW (1995) Development of delivery systems for introducing<br />

endophytic bacteria into cotton. Biocontrol Sci Technol 5:407–416<br />

Neiendam-Nielson M, Sörensen J, Fels J, Pedersen HC (1998) Secondary metabolite- and<br />

endochitinase-dependent antagonism toward plant-pathogenic microfungi of Pseudomonas<br />

fluorescens isolates from sugar beet rhizosphere. Appl Environ Microbiol<br />

64:3563–3569<br />

Nejad P, Johnson PA (2000) Endophytic bacteria induce growth promotion and wilt disease<br />

suppression in oilseed rape and tomato. Biol Control 18:208–215<br />

O’Sullivan DJ, O’Gara F (1992) Traits of fluorescent Pseudomonas spp. involved in suppression<br />

of plant root pathogens. Microbiol Rev 56:662–676<br />

Pleban S, Ingel F, Chet I (1995) Control of Rhizoctonia solani and Sclerotium rolfsii in the<br />

greenhouse using endophytic Bacillus spp. Eur J Plant Pathol 101:665–672<br />

Pleban S, Chernin L, Chet I (1997) Chitinolytic activity of an endophytic strain of Bacillus.<br />

Lett Appl Microbiol 25:284–288<br />

Raaijmakers JM, Weller DM, Thomoshow LS (1997) Frequency of antibiotic-producing<br />

Pseudomonas spp. in natural environments. Appl Environ Microbiol 63:881–887<br />

Reiter B, Pfeifer U, Schwab H, Sessitsch A (2002) Response of endophytic bacterial communities<br />

in potato plants to infection with Erwinia carotovora subsp. atroseptica.Appl<br />

Environ Microbiol 68:2261–2268<br />

Salles JF, De Souza FA, van Elsas JD (2001) Molecular method to assess the diversity of<br />

Burkholderia species in environmental samples. Appl Environ Microbiol 68:1595–1603<br />

Sessitsch A, Reiter B, Pfeifer U, Wilhelm E (2002) Cultivation-independent population<br />

analysis of bacterial endophytes in three potato varieties based on eubacterial and<br />

Actinomycetes-specific PCR of 16S rRNA genes. FEMS Microbiol Ecol 39:23–32<br />

Sessitsch A, Reiter B, Berg G (2004) Endophytic bacterial communities of field-grown potato<br />

plants and their plant growth-promoting abilities. Can J Microbiol 50:239–249<br />

Sharma VK, Nowak J (1998) Enhancement of Verticillium wilt resistance in tomato transplants<br />

by in vitro co-culture of seedlings with a plant growth promoting rhizobacterium<br />

(Pseudomonas sp.) strain PsJN. Can J Microbiol 44:528–536<br />

Siciliano DS, Forin N, Mihoc A, Wisse G, Labell S, Beaumier D, Ouellette D, Roy R, Whyte LG,<br />

Banks MK, Schwab P, Lee K, Greer CW (2001) Selection of specific endophytic bacterial<br />

genotypes by plants in response to soil contamination. Appl Environ Microbiol 67:2469–<br />

2475<br />

Smalla K (2004) Culture-independent microbiology. In: Bull AT (ed) Microbial diversity and<br />

bioprospecting. ASM, Washington DC, pp 88–99<br />

Sturz AV, Christie BR, Matheson BG (1997) Associations of bacterial endophyte populations<br />

from red clover and potato crops with potential for beneficial allelopathy. Can J Microbiol<br />

44:162–167<br />

Sturz AV, Christie BR, Matheson BG, Arsenault WJ, Buchanan NA (1999) Endophytic bacterial<br />

communities in the periderm of potato tubers and their potential to improve<br />

resistance to soilborne plant pathogens. Plant Pathol 48:360–369<br />

Sturz AV, Christie BR, Nowak J (2000) Bacterial endophytes: potential role in developing<br />

sustainable systems of crop production. Crit Rev Plant Sci 19:1–30


4 Control of Plant Pathogenic Fungi with Bacterial Endophytes 69<br />

Tjamos EC, Tsitsigiannis DI, Tjamos SE, Antoniou P, Katinakis P (2004) Selection and<br />

screening of endorhizosphere bacteria from solarised soils as biocontrol agents against<br />

Verticillium dahliae of solanaceous hosts. Eur J Plant Pathol 110:35–44<br />

Weller DM (1988) Biological control of soilborne plant pathogens in the rhizosphere with<br />

bacteria. Annu Rev Phytopathol 26:379–407<br />

Whipps JM (1997) Ecological considerations involved in commercial development of biological<br />

control agents for soil-borne diseases. In: Van Elsas JD, Trevors JT, Wellington EMH<br />

(eds) Modern soil microbiology. Dekker, New York, pp 525–545<br />

Whipps J (2001) Microbial interactions and biocontrol in the rhizosphere. J Exp Bot<br />

52:487–511<br />

Widmer F, Seidler RJ, Gillevet PM, Watrud LS, Di Giovanni GD (1998) A highly selective<br />

PCR protocol for detecting 16S rRNA genes of the genus Pseudomonas (sensu stricto)<br />

in environmental samples. Appl Environ Microbiol 64:2545–2553<br />

Zhou J (2003) Microarrays for bacterial detection and microbial community analysis. Curr<br />

Opin Microbiol 6:288–294<br />

ZinnielDK,LambrechtP,HarrisNB,FengZ,KuczmarskiD,HigleyP,IshimaruCA,Arunakumari<br />

A, Barletta RG, Vidaver AK (2002) Isolation and characterization of endophytic<br />

colonizing bacteria from agronomic crops and prairie plants. Appl Environ Microbiol<br />

68:2198–2208


5<br />

Role of Proteins Secreted by Rhizobia<br />

in Symbiotic Interactions<br />

with Leguminous Roots<br />

Maged M. Saad, William J. Broughton, William J. Deakin<br />

5.1<br />

Introduction<br />

Rhizobia are Gram-negative soil inhabitants with the ability to induce<br />

the formation of highly specialised organs called nodules on the roots or<br />

stems of leguminous plants. Some rhizobial species provoke nodule formation<br />

on a limited number of legume genera and are said to have narrow<br />

host ranges, e.g. Rhizobium meliloti, which nodulates only three genera<br />

of legumes. Other (broad host-range) rhizobia provoke the formation of<br />

nodules on many different legumes, e.g. Rhizobium sp. NGR234 (hereafter<br />

called NGR234), which nodulates more than 112 genera of legumes as well<br />

as the non-legume Parasponia andersonii (Pueppke and Broughton 1999;<br />

Trinick 1980).<br />

To form root nodules, legume roots undergo several new developmental<br />

changes. Initially, rhizobia attach to root hairs, causing deformation and<br />

then curling of the root hair. Rhizobia invade the root through newly formed<br />

tubular structures, called infection threads, which grow toward the root cortex.<br />

During invasion, rhizobia cause the induction of division of cortical<br />

cells, thus forming nodule primordia (Relić et al. 1994). Infection threads<br />

travel inter- and intra-cellularly toward the primordia. Wall-degrading enzymes<br />

help the passage of infection threads from cell to cell (van Spronsen<br />

et al. 1994). Rhizobia are released from the infection threads into the cytoplasm<br />

of host cells by a process resembling endocytosis (Stacey et al. 1991).<br />

Extensive cell division in the primordia leads to functional nodules containing<br />

rhizobia, in which the bacteria differentiate into their endosymbiotic<br />

form, known as the bacteroids (Franssen et al. 1992). Bacteroids, together<br />

with the surrounding plant-derived peribacteroid membrane (PBM) are<br />

Maged M. Saad: Université de Genève, Laboratoire de Biologie Moléculaire des Plantes<br />

Supérieures, Sciences III, 30 Quai Ernest-Ansermet, 1211 Genève 4, Switzerland<br />

William J. Broughton: Université de Genève, Laboratoire de Biologie Moléculaire des Plantes<br />

Supérieures, Sciences III, 30 Quai Ernest-Ansermet, 1211 Genève 4, Switzerland, E-mail:<br />

william.broughton@bioveg.unige.ch<br />

William J. Deakin: Université de Genève, Laboratoire de Biologie Moléculaire des Plantes<br />

Supérieures, Sciences III, 30 Quai Ernest-Ansermet, 1211 Genève 4, Switzerland<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


72 M.M. Saad et al.<br />

called symbiosomes. At this stage, the bacteria synthesise nitrogenase, an<br />

enzyme complex that catalyses the reduction of atmospheric nitrogen to<br />

ammonia, which is subsequently assimilated into amino acids. In return,<br />

the plant reduces carbon dioxide to sugars during photosynthesis and<br />

translocates these compounds to the roots, where the bacteria use them<br />

as an energy source (Atkins et al. 1982; Fisher and Long 1992). Depending<br />

upon the plant species, at least two types of nodules are formed. Indeterminate<br />

nodules have a persistent meristem that grows continuously giving<br />

the nodules an elongated shape. Determinant nodules lack the persistent<br />

meristem, and are round in shape as the meristematic activity is limited to<br />

the early stages of nodule development.<br />

Coordination of this complex developmental programme requires the<br />

exchange of many signals between the two symbiotic partners and it is<br />

the response to (and synchronisation of) these signals that controls nodule<br />

formation. Amongst the first signal molecules are phenolic compounds,<br />

mainly flavonoids that are secreted by roots into the rhizosphere. The rhizobial<br />

protein NodD functions first as an environmental sensor of these<br />

phenolics, and later as a transcriptional activator of a series of rhizobial<br />

genes that encode proteins responsible for the production of rhizobial<br />

signals. NodD proteins belong to the LysR family of transcriptional regulators,<br />

which have the ability to bind to specific, highly conserved DNA<br />

sequences (nod-boxes) present in the promoter regions of many nodulation<br />

genes/loci (Perret et al. 2000). The number of nod-boxes varies in different<br />

rhizobia. For example, in NGR234 there are at least 19 nod-boxes that help<br />

regulate transcription of genes involved in a range of different signalling<br />

compounds (Kobayashi et al. 2004). The first rhizobial signal molecules<br />

to be synthesised and secreted are encoded by the nodulation genes (nod,<br />

noe,andnol), which are responsible for the synthesis of host-specific lipochito-oligosaccharide<br />

molecules called Nod-factors. Nod-factors provoke<br />

deformation and curling of the root hair and allow rhizobia to enter roots<br />

through infection-threads (Relić et al. 1993, 1994; D’Haeze et al. 1998).<br />

All invasive rhizobia produce Nod-factors, and the addition of purified<br />

Nod-factors alone causes root-hair deformation and cortical-cell division<br />

(Downie 1998; Perret et al. 2000). Nod-factors consist of a backbone of three<br />

to six β-1.4-linked N-acetyl-d-glucosamine residues. A fatty-acid chain of<br />

variable length and structure (depending on the Rhizobium species) is attached<br />

at the C-2 position of the non-reducing sugar residue. Synthesis<br />

of the backbone is brought about by the products of the nodABC genes,<br />

known as the core enzymes, as they are found in all rhizobia. NodB and<br />

NodC are responsible for the synthesis of the backbone while NodA is<br />

an acyl-transferase, which adds the fatty-acid side chain. Nod-factors are<br />

further modified by the action of other Nod enzymes that add various<br />

chemical groups to the backbone (Hanin et al. 1999; Broughton et al. 2000).


5 Protein secretion by Rhizobia 73<br />

The development of the infection threads needs other sets of rhizobial<br />

signals, including surface polysaccharides (SPS) as well as secreted proteins.<br />

SPSs include extracellular polysaccharides (EPS), capsular polysaccharides<br />

(CPS), lipo-polysaccharides (LPS) as well as cyclic β-glucans. Most of these<br />

polysaccharides function during infection thread development, where they<br />

possibly help suppress plant defence reactions. SPSs have highly diverse<br />

structures and may contribute to rhizobial host-range (for reviews, see<br />

Broughton et al. 2000; Perret et al. 2000). In this chapter we will concentrate<br />

on protein secretion by rhizobia, which are thought to play a role in<br />

the infection process leading to nodule formation and, thus, in successful<br />

nitrogen fixation.<br />

5.2<br />

Bacterial Protein Secretion Systems<br />

Gram-negative bacteria possess at least four systems (type I–type IV) to<br />

secrete proteins into the external environment (see Fig. 5.1) (Thanassi<br />

and Hultgren 2000). Examples of all four systems have been found in<br />

rhizobia. Type II secretion is said to be sec-dependent as it requires the<br />

sec system to export proteins into the periplasm prior to secretion across<br />

the outer membrane. Proteins secreted by the type II system thus possess<br />

a classical amino-terminal hydrophobic signal sequence, which is cleaved<br />

during sec-export. In contrast, type I and type III secretion systems are secindependent;<br />

proteins are secreted across the bacterial inner- and outermembranes<br />

in a single step process. Such proteins do not possess cleavable<br />

amino-terminal signal sequences. A number of proteins secreted by type I<br />

or type III secretion systems are able to directly affect nodulation in a variety<br />

of legume-rhizobia associations.<br />

5.2.1<br />

Type I Secretion Systems<br />

Many Gram-negative bacteria utilise type I [or ATP-binding cassette (ABC)]<br />

secretion systems (T1SS). Generally, the substrates of ABC transporters are<br />

toxins, proteases or lipases. All T1SS secreted proteins possess a carboxyterminal<br />

secretion signal of approximately 60 amino acids, which is not<br />

cleaved during export. The secretion machine consists of multimers of<br />

three proteins: an inner membrane exporter, an outer membrane pore, and<br />

an inner-membrane anchored protein that spans the periplasm linking the<br />

proteins found in the inner and outer membranes. Proteins can thus be<br />

directly secreted from the cytoplasm to the external environment without


74 M.M. Saad et al.


5 Protein secretion by Rhizobia 75<br />

◮ Fig.5.1.Diagrammatic scheme describing four different types of protein secretion systems<br />

in Gram-negative bacteria. OM Bacterial outer membrane, IM bacterial inner membrane,<br />

PP periplasmic space, PM host plasma membrane. A number of proteins are involved in<br />

the assembling of the different secretion apparatus (indicated by spheres and ellipsoids).<br />

T2SS (type II secretion system) and T4SS (type IV secretion system) are Sec-dependent,<br />

thus proteins to be exported from the bacterial cell are first transported to the periplasm by<br />

the Sec system. The direction of this two-step secretion is shown by the black arrow, from<br />

the periplasm to the external environment. T4SS can also export other protein substrates<br />

directly from the cytosol, which does not require the Sec system. T1SS (type I secretion<br />

system) and T3SS (type III secretion system) are Sec-independent, transporting proteins<br />

directly from the bacterial cytoplasm to the external environment or into eukaryotic cells<br />

(for T3SS). Extracellular appendages are known to be components of several T3SS and<br />

T4SS. Secreted proteins are represented by black circles if they are exported from the<br />

periplasm (Sec-dependent) or black squares if they originate in the bacterial cytoplasm<br />

(Sec-independent)<br />

the formation of periplasmic intermediates (Hueck 1998; Thanassi and<br />

Hultgren 2000).<br />

T1SSs Involved in Symbiosis<br />

T1SSs have been found in a number of rhizobia, including Rhizobium<br />

leguminosarum bv. viciae and Rhizobium sp. BR816 (van Rhijn et al. 1996).<br />

NodO was first identified as a secreted protein in R. leguminosarum bv.<br />

viciae by de Maagd et al. (1989). The nodO gene is flavonoid inducible<br />

in a nod-box- and nodD-dependent manner (de Maagd et al. 1989; van<br />

Rhijn et al. 1996). NodO exists as a dimer in its native form, with an<br />

approximate molecular mass of 67 kDa (Sutton et al. 1994, 1996). The<br />

amino acid sequence of NodO contains putative repeated Ca 2+ -binding<br />

domains in the amino-terminal region, and has homology to a number<br />

of pore-forming bacterial toxins known as RTX proteins (Economou et<br />

al. 1990). Studies in vitro have shown that purified NodO forms cationselective<br />

pores in plasma-membrane lipid bilayers (Sutton et al. 1994). By<br />

analogy, NodO may thus form a pore in the root cell membrane, causing an<br />

influx of Ca 2+ ,whichcouldactasasecondmessengertherebystimulating<br />

the cytoskeletal changes required for infection thread growth.<br />

Mutation of the nodO gene has little effect on nodulation (Downie and<br />

Surin 1990), although double mutants of nodO and those involved in Nodfactor<br />

synthesis display clear Nod − phenotypes. This was unexpected as<br />

NodO is not involved in the biosynthesis or export of Nod-factors (Spaink<br />

et al. 1991; Sutton et al. 1994). Yet on Pisum sativum and Vicia sativa<br />

for example, NodO and a functional nodE are necessary for nodulation<br />

(Downie and Surin 1990; Economou et al. 1994). As infection threads appear<br />

to abort in the nodO/nodE doublemutant,itispossiblethatNodE(an


76 M.M. Saad et al.<br />

α-keto-acyl-synthase) is also involved in acylation of some components of<br />

the infection thread. Furthermore, mobilising nodO into different rhizobia<br />

extends the host range of the trans-conjugant (e.g. nodO into a nodE<br />

mutant of R. leguminosarum bv. trifolii allows transconjugants to nodulate<br />

V. sativa; Economou et al. 1994), thus perhaps pointing to a host-specific<br />

role for NodO. Another example of nodO complementing a defect in Nodfactor<br />

synthesis was shown when a nodO homologue of Rhizobium sp.<br />

strain BR816 was used to complement a nodS mutant of NGR234 for the<br />

nodulation of Leucaena leucocephala (van Rhijn et al. 1996). Again, NodO<br />

and NodS have distinct biochemical functions: NodS is a N-methyl transferase<br />

that methylates Nod-factors (Geelen et al. 1995; Jabbouri et al. 1995).<br />

The BR816 nodO gene was also shown to suppress the nodulation defect<br />

of the nodU mutants of NGR234 and R. tropici CIAT899 on L. leucocephala<br />

and of the nodE mutant of R. leguminosarum bv. trifolii ANU842 on white<br />

clover (Vlassak et al. 1998).<br />

Based on the observation that over-expression of nodO rescues nodulation<br />

by the multiple mutant nodFEMNTLO (of R. leguminosarum bv.<br />

viciae), Walker and Downie (2000) proposed a role for NodO in the complementation<br />

of Nod-factor defects. This mutant produces Nod-factors that<br />

are devoid of decorations and results in abortion of infection-thread development<br />

in Vicia sativa. They suggested that NodO stimulates ion flow<br />

across the cell membrane, thereby amplifying a weaker-than-normal signal<br />

transmitted by the undecorated version of Nod-factor.<br />

NodO is not the only T1SS Protein Involved in Symbiosis<br />

Although the phenotype of a R. leguminosarum nodO mutant was Fix +<br />

on P. sativum, inactivation of the type I system that secretes NodO results<br />

in Fix − nodules (Economou et al. 1994). It is thus possible that the T1SS<br />

is capable of secreting other proteins that play important roles in nodulation.<br />

Two such proteins were identified as PlyA and PlyB, two similar<br />

enzymes that function as extra-cellular glycanases, which are involved in<br />

processing rhizobial EPS (Finnie et al. 1998). ExpEI is secreted in a type<br />

I-dependent fashion by Rhizobium meliloti. Like the PlyAB proteins of<br />

R. leguminosarum, ExpEI is thought to be involved in the extra-cellular<br />

processing of EPS (Becker et al. 1997; Moreira et al. 2000). Thus it seems<br />

that proteins secreted via T1SSs in rhizobia play indirect roles in symbiosis,<br />

perhaps by amplifying or modifying other signal molecules. In this<br />

way they could increase ion flux following pore formation in plasma membranes,<br />

as proposed for NodO, or by augmenting the amounts of the active<br />

forms of EPS, as is possibly the case with ExpE1, and the PlyAB enzymes.


5 Protein secretion by Rhizobia 77<br />

5.2.2<br />

Type II Secretion Systems<br />

A wide variety of Gram-negative bacteria utilise type II secretion systems<br />

(T2SSs) as a stepwise process to export proteins from the periplasm across<br />

the outer membrane. The amino-terminal signal peptides of the secreted<br />

proteins are first recognised and then translocated by a sec-dependent<br />

mechanism through the inner membrane. The signal peptide is cleaved,<br />

releasing the protein into the periplasm (Pugsley 1993; Sandkvist 2001). The<br />

T2SS is also called the general secretory pathway (GSP), and is responsible<br />

for the secretion of a large variety of degradative enzymes and toxins. T2SSs<br />

are composed of a core of between 12 and 15 proteins (Fig. 5.1), not all of<br />

which are present in every T2SS, as some appear to be dispensable for<br />

secretion (Sandkvist 2001). The core proteins are thought to form a multiprotein<br />

complex, spanning the periplasmic compartment that is specifically<br />

required for the translocation of any secreted proteins across the outer<br />

membrane (Sandkvist 2001; Peabody et al. 2003). In rhizobia, there is no<br />

clear evidence that any T2SSs play a role in symbiosis or nodule formation.<br />

Interestingly, the type II secretion machine shares many features with<br />

the type IV pilus biogenesis system found in many Rhizobium species, e.g.<br />

Bradyrhizobium japonicum USDA110, Mesorhizobium loti MAFF303099,<br />

R. meliloti and NGR234 (Kaneko et al. 2000, 2002; Galibert et al. 2001; Streit<br />

et al. 2004). Type IV pili are found on the surface of many Gram-negative<br />

bacteria, where they play an important role in bacterial adhesion to host<br />

cells, bio-film formation and conjugative DNA transfer (Wolfgang et al.<br />

2000). Nitrogen fixing bacteria of the genus Azoarcus utilise type IV pili to<br />

colonise grasses (Dörr et al. 1998). It remains to be seen whether rhizobia<br />

usetypeIVpiliduringthesymbioticinteractionwithlegumes.<br />

5.2.3<br />

Type III Secretion Systems<br />

Type III secretion systems (T3SSs) are characteristic of pathogenic Gramnegative<br />

bacteria, where their function is to inject proteins into the cytoplasm<br />

of eukaryotic cells, so facilitating the onset of disease. The T3SS<br />

is composed of a complex of about 20 proteins that spans both bacterial<br />

membranes (Fig. 5.1). Ten of these proteins are highly conserved in all<br />

T3SS-possessing bacteria and even show similarities to components of the<br />

flagella assembly apparatus, from which the pathogenic T3SS are thought<br />

to have evolved (Hueck 1998). Proteins that are secreted by T3SSs can be<br />

separated into four classes based on their functions. Some of them polymerise<br />

into extra-cellular components of the secretion apparatus forming


78 M.M. Saad et al.<br />

pili. There are also effector proteins that are actually injected into the<br />

cytosol of host cells, which then modulate cellular functions of the host<br />

by interfering with signalling cascades or disrupting the cytoskeleton. The<br />

third class of secreted proteins is termed translocators, and they polymerise<br />

to form a pore in the membrane of eukaryotic cells that allows the effectors<br />

to pass into the cells (Hueck 1998; Feldman and Cornelis 2003). Finally, in<br />

certain T3SS-possessing bacteria, secreted regulatory proteins that control<br />

cell contact-dependent secretion have also been identified (He 1998; Hueck<br />

1998). Proteins secreted by a T3SS do not require the sec system for their<br />

transit from the bacterial cytoplasm to the eukaryotic cell, although the sec<br />

pathway might be required for assembly of the type III secretion apparatus<br />

within the bacterial membranes. (Several components of the apparatus<br />

carry sec-characteristic amino-terminal signal sequences; Hueck 1998).<br />

Identification and Function of T3SSs in Rhizobia<br />

Given their importance in pathogenicity, it was a surprise to find T3SSs in<br />

symbiotic rhizobia. A complete T3SS was first identified in Rhizobium sp.<br />

NGR234 (Freiberg et al. 1997). All ten genes encoding the conserved components<br />

of T3SSs were found and named rhc (Rhizobium conserved) but<br />

thesamefinalletterasusedtodescribepathogenicT3SSgeneswasmaintained<br />

(Viprey et al. 1998). Sequencing other large replicons of B. japonicum<br />

USDA110 (Göttfert et al. 2001; Kaneko et al. 2002) and M. loti MAFF303099<br />

(Kaneko et al. 2000) also revealed the existence of loci encoding T3SSs.<br />

T3SSs are not ubiquitous in rhizobia, however, as no such system was<br />

found within the completed genome of R. meliloti (Galibert et al. 2001).<br />

Random mutagenesis also identified mutants of R. fredii strains that were<br />

affected in nodulation. Subsequent analysis of the insertion sites showed<br />

them to be within T3SSs (Marie et al. 2001; Krishnan et al. 2003). In fact,<br />

mutagenesis of T3SSs of rhizobia proved that although they are not absolutely<br />

essential for nodulation of all legumes, they have cultivar-specific<br />

effects and thus can be viewed as determinants of rhizobial host-range<br />

(Marie et al. 2001). This is exemplified by studies on the T3SS of NGR234<br />

(Viprey et al. 1998). In NGR234, the T3SS genes are grouped within a 30-kb<br />

region of the symbiotic plasmid pNGR234a (Freiberg et al. 1997). Knockout<br />

mutants in the T3SS machine of NGR234 cause three host-dependent<br />

effects. As compared to the wild-type rhizobia, there can be a dramatic impairment<br />

of nodule development e.g. on Tephrosia vogelii, resulting in the<br />

formation of a majority of non-fixing pseudo-nodules. Second, a dramatic<br />

enhancement of nodulation is often seen, as with Crotalaria juncea and<br />

Pachyrhizus tuberosus, while a third group of legumes seems to be unaffected<br />

by the presence/absence of a functional T3SS (e.g. Lotus japonicus<br />

and Vigna unguiculata; Marie et al. 2003; Viprey et al. 1998).


5 Protein secretion by Rhizobia 79<br />

Regulation of Rhizobial T3SSs<br />

In rhizobia, transcription of the T3SS-related genes requires the presence<br />

of flavonoids and two bacterial regulatory elements: NodD1 and TtsI<br />

(Krishnan et al. 1995; Viprey et al. 1998; Krause et al. 2002). TtsI shares<br />

homology with transcriptional activators of the two-component sensorregulator<br />

family (Viprey et al. 1998). The gene encoding TtsI is found within<br />

the T3SS loci of rhizobia and is preceded by a nod-box. It has been shown<br />

that, after flavonoid activation, NodD1 induces ttsI transcription, which in<br />

turn activates genes within T3SS loci (Viprey et al. 1998; Kobayashi et al.<br />

2004). Induction of the T3SS genes occurs after induction of genes involved<br />

in Nod-factor synthesis, implying that the T3SS functions after Nod-factors<br />

in the symbiosis. A conserved promoter motif called the tts-box has been<br />

identified upstream of most T3SS regulons (Krause et al. 2002). Although<br />

it has not been demonstrated experimentally, it is thought that TtsI may<br />

bind to tts-boxes to induce transcription of the downstream genes. Transcriptional<br />

studies suggest that the T3SS of NGR234 does not function<br />

throughout the symbiosis, as the majority of T3SS gene transcripts could<br />

not be detected in nodules of V. unguiculata and Cajanus cajan (Perret et<br />

al. 1999).<br />

Functions of Proteins Secreted via Rhizobial T3SSs<br />

Protein secretion by T3SSs of rhizobia has been demonstrated in vitro<br />

for R. fredii USDA257 and NGR234. Proteins secreted in a T3SS-dependent<br />

manner are called Nops (nodulation outer proteins) (Marie et al. 2001). Five<br />

Nops are known to be secreted by USDA257 (Krishnan and Pueppke 1993)<br />

and at least eight by NGR234 (Marie et al. 2003). Several of these Nops have<br />

been identified and subsequently characterised. Functional studies have<br />

shown that rhizobial Nops can be placed into three of the general classes<br />

of type-III-secreted proteins. Figure 5.2 summarises the role of the Nops<br />

within a rhizobial T3SS.<br />

External Components of the Machinery<br />

Some type-III-secreted proteins of phyto-pathogenic bacteria have been<br />

shown to polymerise and form pili. These pili serve to connect the bacterium<br />

to the plant cell and allow proteins to pass through the hollow pili.<br />

As protein secretion is abolished in mutants of pili-genes, their phenotype<br />

resembles that produced by knock-outs of the T3SS machine (He and<br />

Jin 2003). In the presence of flavonoids, USDA257 produces extra-cellular<br />

appendages (pili), and requires a functional T3SS (Krishnan et al. 2003).<br />

Preliminary results indicate that NGR234 also produces pilus-like structures<br />

on its surface in a flavonoid and T3SS-dependent manner. These appendages<br />

were purified and shown to be composed predominantly of NopA


80 M.M. Saad et al.<br />

Fig.5.2. Hypothetical structure of the T3SS of Rhizobium sp. NGR234, adapted from Viprey<br />

et al. (1998) and Bartsev et al. (2004b). The conserved components (Rhc proteins) of the<br />

T3SS form a channel through the bacterial inner and outer membranes. The known roles<br />

of the Nops are also illustrated. NopA and NopB are thought to be the major components<br />

of a T3SS-dependent pilus that links the rhizobial cell to the plant cell. Nops are secreted<br />

through the pilus and can thus cross the plant cell wall. NopX may polymerise to form<br />

a pore in the plant root cell plasma membrane and, finally, NopL and NopP are possible<br />

effector proteins that function within the plant root cell. NGR234 probably secretes many<br />

other effector proteins<br />

(W.J. Deakin, unpublished data), the smallest protein secreted by NGR234<br />

(Marie et al. 2003). Furthermore, homologues of nopA have been identified<br />

in all T3SS-possessing rhizobia, suggesting that NopA is an essential<br />

component of the secretion machinery. Mutation of nopA also blocks the secretion<br />

of all the other Nops. This phenotype resembles that of mutations in<br />

other genes that encode external components of the T3SS. Although NopA<br />

is the major component of the rhizobial T3SS-pili, mutations in other genes<br />

that encode Nops also block Nop secretion, suggestion that other, minor<br />

components of the external secretion apparatus might exist (unpublished<br />

data).<br />

Translocators<br />

NopX, one of the first Nops to be identified in NGR234 (Viprey et al. 1998),<br />

has significant homology to a number of proteins secreted by T3SSs of


5 Protein secretion by Rhizobia 81<br />

phytopathogens. Perhaps the best studied of these is HrpF of Xanthomonas<br />

campestris pv. vesicatoria (Huguet and Bonas 1997). HrpF is secreted in<br />

a T3SS-dependent manner and may function as a translocator of the effectors<br />

proteins into host cells (Rossier et al. 2000). Indeed, HrpF has<br />

been shown to form pores in lipid bilayers (Büttner et al. 2002) We have<br />

thus proposed that NopX could perform the role of translocator for rhizobial<br />

T3SSs (Marie et al. 2003). NopX of USDA257 has been localised to<br />

the infection threads, where it could play a role in the infection thread<br />

growth (Krishnan 2002). Although nopX homologues are present in most<br />

T3SS-possessing rhizobia, there does not appear to be a homologue in<br />

B. japonicum USDA110 (Krause et al. 2002). This is extremely puzzling as<br />

the translocon is thought to be essential for the transport of T3SS proteins<br />

into eukaryotic cells.<br />

Effectors<br />

This group of secreted proteins are thought to function within the root<br />

cells. So far, only one example of a rhizobial T3SSs effector has been identified<br />

(NopL), although it is suspected that there could be many more.<br />

Homologues of nopL are found only in T3SS-possessing rhizobia, although<br />

M. loti MAFF303099 does not appear to have a copy. Mutations in genes<br />

encoding effector proteins do not affect the secretion of any other T3SS proteins,<br />

and this was shown to be the case for a mutation in nopL of NGR234<br />

(Marie et al. 2003). A nopL mutant has a similar nodulation phenotype<br />

toNGR234onthemajorityoflegumestested.Thisisanothercharacteristicofeffectorproteinsofphytopathogens,fortherearemanyofthem<br />

and they are thought to be redundant in function. NopL is important for<br />

the efficient nodulation of Flemingia congesta (Marie et al. 2003), suggesting<br />

that it is a rhizobial “virulence factor” for this plant. Functional<br />

characterisation of NopL revealed that it can be phosphorylated by plant<br />

kinases (Bartsev et al. 2003). Furthermore, expression of nopL in Nicotiana<br />

tabacum inhibited this plant’s ability to accumulate pathogenesisrelated<br />

(PR)-proteins in response to pathogen attack. We thus suggest that<br />

NopL could suppress root-cell defence responses by disrupting the intracellular<br />

signalling cascades required for activation of PR-genes (Bartsev et<br />

al. 2004a).<br />

Sequence analyses of NGR234 and USDA110 revealed the presence of<br />

geneshomologoustosecretedeffectorproteinsfromotherT3SS-possessing<br />

pathogenic bacteria. The proteins encoded by these rhizobial genes are thus<br />

good candidates for secretion in a T3SS-dependent manner, and possibly<br />

function as effectors within legume root cells (Marie et al. 2001, Krause<br />

et al. 2002). These proteins have been studied extensively in pathogenic<br />

bacteria, where they act to suppress host defence responses, and it will be


82 M.M. Saad et al.<br />

interesting to determine whether their rhizobial counterparts function in<br />

a similar manner.<br />

The Role of Rhizobial T3SSs<br />

At this stage, explanations for the roles of T3SSs in rhizobia can only be<br />

hypothetical. It is thought that the T3SS functions during infection thread<br />

development and perhaps while the rhizobia are being released from infection<br />

threads into cortical cells. Similarities between attempts by rhizobia<br />

to colonise roots and the attack by pathogens of other plant cells are obvious.<br />

Undoubtedly, legumes mount defences against rhizobia. NGR234<br />

and similar rhizobia may have acquired a T3SS to suppress such defence<br />

responses, and effectors like NopL are the inter-cellular messengers in this<br />

process. Legumes, like most other plants have probably evolved sophisticated<br />

methods for detecting T3SS-containing pathogens. To some plants,<br />

NGR234 (and similar rhizobia) would reveal themselves as “pathogens”<br />

provoking defence-responses that block nodulation (e.g. P. tuberosus). In<br />

this scenario, other legumes would have evolved immunity to T3SS proteins<br />

(L. japonicus, V. unguiculata), while still others would positively welcome<br />

them (e.g. T. vogeli). Furthermore, it is possible that rather than responding<br />

in one of three ways to T3SS-proteins, a continuum of responses exists,<br />

some too slight to be detected. Other possibilities include that the T3SS is<br />

for some reason not activated by certain plants, or that some plants possess<br />

other signalling systems that over-ride or complement the T3SS.<br />

5.2.4<br />

Type IV Secretion Systems<br />

Type IV secretion systems (T4SSs) were initially defined on the basis of<br />

the homologies between components of three different macromolecular<br />

complexes: the Agrobacterium tumefaciens T-DNA transfer system that is<br />

required for exporting oncogenic T-DNA to susceptible plant cells; the conjugal<br />

transfer (Tra) system of the conjugative IncN plasmid pKM101; and<br />

the Bordetella pertussis toxin exporter (Ptl) machine (Winans et al. 1996;<br />

Christie 1997). Like T2SSs, T4SSs use a stepwise process to translocate<br />

macromolecular substrates first across the inner membrane, prior to transport<br />

across the cell envelope (Christie 2001). Some symbiotic nitrogenfixing<br />

bacteria also possess genes that could encode a T4SS e.g. R. etli,<br />

M. loti strain R7A and R. meliloti (Galibert et al. 2001; Sullivan et al. 2002;<br />

Gonzalez et al. 2003). It is interesting to note that in M. loti R7A the location<br />

of the genes that may encode a T4SS is exactly at the T3SS locus of<br />

M. loti MAFF303099 strain, although each strain possesses only one type of<br />

secretion machine. The role of T4SSs in symbiosis is not known, but there


5 Protein secretion by Rhizobia 83<br />

is a suggestion that it could affect the nodulation process, as a putative<br />

nod-box is located in the promoter region of one of the genes of the R7A<br />

T4SS (Sullivan et al. 2002).<br />

5.3<br />

Conclusions<br />

Successful symbiotic associations between rhizobia and legumes require<br />

the exchange of many signal molecules. Both partners secrete these signals<br />

and it is the timing of their emission and perception as well as the quantity<br />

that are probably important. Symbiotic harmony depends on the precise<br />

meshing of these signals. Plant flavonoids are the first important group of<br />

signalling molecules and they act as inducers of nodulation genes (nod,<br />

noe and nol) (Broughton et al. 2000; Perret et al. 2000). The regulatory<br />

networks of flavonoids, the NodD family of transcriptional activators, and<br />

their conserved promoter sequences (nod-boxes) guarantee the timing of<br />

expression of downstream genes that are responsible for the synthesis of<br />

diffusible lipo-chito-oligosaccharidic Nod-factors – early symbiotic “master<br />

keys”. Once the legume “doors” have been opened to allow rhizobia in,<br />

different morphological and cytological changes in the roots occur. Nodfactors<br />

play only secondary roles in the later steps of invasion, at which<br />

time other signal molecules occupy centre stage. Bacterial SPSs and secreted<br />

proteins contribute to the infection process, where they assist in<br />

infection thread development within the root hair, and help modify hostdefence<br />

mechanisms.<br />

Acknowledgements. We thank D. Gerber for general support and encouragement.<br />

Research in LBMPS is financed by the Founds National Suisse<br />

de la recherché Scientifique (Project 31-63893.00) and the Université de<br />

Genève.<br />

<strong>References</strong><br />

Atkins CA, Ritchie A, Rowe PB, McCairns E, Sauer D (1982) De novo purine synthesis in<br />

nitrogen-fixing nodules of cowpea (Vigna unguiculata [L.] Walp.) and soybean (Glycine<br />

max [L.] Merr.). Plant Physiol 70:55–60<br />

Bartsev AV, Boukli NM, Deakin WJ, Staehelin C, Broughton WJ (2003) Purification and<br />

phosphorylation of the effector protein NopL from Rhizobium sp. NGR234. FEBS Lett<br />

554:271–274<br />

Bartsev AV, Deakin WJ, Boukli NM, McAlvin CB, Stacey G, Malnöe P, Broughton WJ, Staehelin<br />

C (2004a) NopL, an effector protein of Rhizobium sp. NGR234, thwarts activation<br />

of plant defense reactions. Plant Physiol 134:871–879


84 M.M. Saad et al.<br />

Bartsev AV, Kobayashi H, Broughton WJ (2004b) Rhizobial signals convert pathogens to<br />

symbionts at the legume interface. In: Gillings M, Holmes A (eds) Plant microbiology,<br />

Bios Scientific, Oxfordshire, UK, pp 19–28<br />

Becker A, Ruberg S, Kuster H, Roxlau AA, Keller M, Ivashina T, Cheng HP, Walker GC,<br />

Pühler A (1997) The 32-kilobase exp gene cluster of Rhizobium meliloti directing the<br />

biosynthesis of galactoglucan: genetic organization and properties of the encoded gene<br />

products. J Bacteriol 179:1375–1384<br />

Broughton WJ, Jabbouri S, Perret X (2000) Keys to symbiotic harmony. J Bacteriol 182:5641–<br />

5652<br />

Büttner D, Nennstiel D, Klusener B, Bonas U (2002) Functional analysis of HrpF, a putative<br />

type III translocon protein from Xanthomonas campestris pv. vesicatoria. JBacteriol<br />

184:2389–2398<br />

Christie PJ (1997) Agrobacterium tumefaciens T-complex transport apparatus: a paradigm<br />

for a new family of multifunctional transporters in eubacteria. J Bacteriol 179:3085–3094<br />

Christie PJ (2001) Type IV secretion: intercellular transfer of macromolecules by systems<br />

ancestrally related to conjugation machines. Mol Microbiol 40:294–305<br />

De Maagd RA, Wijfjes AH, Spaink HP, Ruiz-Sainz JE, Wijffelman CA, Okker RJ, Lugtenberg<br />

BJ (1989) nodO, anewnod gene of the Rhizobium leguminosarum biovar viciae<br />

sym plasmid pRL1JI, encodes a secreted protein. J Bacteriol 171:6764–6770<br />

D’Haeze W, Gao M-S, De Rycke R, Van Montagu M, Engler G, Holsters M (1998) Roles<br />

for Azorhizobial Nod factors and surface polysaccharides in intercellular invasion and<br />

nodule penetration, respectively. Mol Plant Microbe Interact 11:999–1008<br />

Downie JA (1998) Functions of rhizobial nodulation genes. In: Spaink HP, Kondorosi A,<br />

Hooykaas JP (eds) The Rhizobiaceae. Kluwer, Dordrecht, pp 387–402<br />

Dörr J, Hurek T, Reinhold-Hurek B (1998) Type IV pili are involved in plant-microbe and<br />

fungus-microbe interactions. Mol Microbiol 30:7–17<br />

Downie JA, Surin BP (1990) Either of two nod gene loci can complement the nodulation<br />

defect of a nod deletion mutant of Rhizobium leguminosarum bv viciae.MolGenGenet<br />

222:81–86<br />

Economou A, Hamilton WD, Johnston AW, Downie JA (1990) The Rhizobium nodulation<br />

gene nodO encodes a Ca 2+ binding protein that is exported without N-terminal cleavage<br />

and is homologous to haemolysin and related proteins. EMBO J 9:349–354<br />

Economou A, Davies AE, Johannes E, Downie JA (1994) The Rhizobium leguminosarum<br />

biovar vicia nodO gene can enable a nodE mutant of Rhizobium leguminosarum biover<br />

trifolii to nodulate vetch. Microbiology 140:2341–2347<br />

Feldman MF, Cornelis GR (2003) The multitalented type III chaperones: all you can do with<br />

15 kDa. FEMS Microbiol Lett 219:151–158<br />

Finnie C, Zorreguieta A, Hartley NM, Downie JA (1998) Characterization of Rhizobium<br />

leguminosarum exopolysaccharide glycanases that are secreted via a type I exporter<br />

and have a novel heptapeptide repeat motif. J Bacteriol 180:1691–1699<br />

Fisher RF, Long SR (1992) Rhizobium-plant signal exchange. Nature 357:655–660<br />

Franssen HJ, Vijn I, Yang WC, Bisseling T (1992) Developmental aspects of the Rhizobiumlegume<br />

symbiosis. Plant Mol Biol 19:89–107<br />

Freiberg C, Fellay R, Bairoch A, Broughton WJ, Rosenthal A, Perret X (1997) Molecular basis<br />

of symbiosis between Rhizobium and legumes. Nature 387:394–401<br />

Galibert F, Finan TM, Long SR, Puhler A, Abola P, Ampe F, Barloy-Hubler F, Barnett MJ,<br />

Becker A, Boistard P, Bothe G, Boutry M, Bowser L, Buhrmester J, Cadieu E, Capela D,<br />

Chain P, Cowie A, Davis RW, Dreano S, Federspiel NA, Fisher RF, Gloux S, Godrie T,<br />

Goffeau A, Golding B, Gouzy J, Gurjal M, Hernandez-Lucas I, Hong A, Huizar L, Hyman<br />

RW, Jones T, Kahn D, Kahn ML, Kalman S, Keating DH, Kiss E, Komp C, Lelaure V,<br />

MasuyD,PalmC,PeckMC,PohlTM,PortetelleD,PurnelleB,RamspergerU,SurzyckiR,


5 Protein secretion by Rhizobia 85<br />

Thebault P, Vandenbol M, Vorholter FJ, Weidner S, Wells DH, Wong K, Yeh KC, Batut J<br />

(2001) The composite genome of the legume symbiont Sinorhizobium meliloti.Science<br />

293:668–672<br />

Geelen D, Goethals K, Van Montagu M, Holsters M (1995) The nodD locus from Azorhizobium<br />

caulinodans is flanked by two repetitive elements. Gene 164:107–111<br />

Gonzalez V, Bustos P, Ramirez-Romero MA, Medrano-Soto A, Salgado H, Hernandez-<br />

GonzalezI,Hernandez-CelisJC,QuinteroV,Moreno-HagelsiebG,GirardL,Rodriguez<br />

O, Flores M, Cevallos MA, Collado-Vides J, Romero D, Davila G (2003) The<br />

mosaic structure of the symbiotic plasmid of Rhizobium etil CFN42 and its relation to<br />

other symbiotic genome compartments. Genome Biol 4:R36<br />

Göttfert M, Röthlisberger S, Kündig C, Beck C, Marty R, Hennecke H (2001) Potential<br />

symbiosis-specific genes uncovered by sequencing a 410-kilobase DNA region of the<br />

Bradyrhizobium japonicum chromosome. J Bacteriol 183:1405–1412<br />

Hanin M, Jabbouri S, Broughton WJ, Fellay R, Quesada-Vincens D (1999) Molecular aspects<br />

of host-specific nodulation. In: Stacey G, Keen NT (eds) Plant-microbe interaction.<br />

American Phytopathological Society, St Paul, MN, pp 1–37<br />

He SY (1998) Type III protein secretion systems in bacterial pathogenic bacteria. Annu Rev<br />

Phytopathol 36:363–392<br />

He SY, Jin Q (2003) The Hrp pilus: learning from flagella. Curr Opin Microbiol 6:15–19<br />

Hueck CJ (1998) Type III protein secretion systems in bacterial pathogens of animals and<br />

plants. Microbiol Mol Biol Rev 62:379–433<br />

Huguet E, Bonas U (1997) hrpF of Xanthomonas campestris pv. vesicatoria encodes an 87kDa<br />

protein with homology to NoIX of Rhizobium fredii. MolPlantMicrobeInteract<br />

10:488–498<br />

Jabbouri S, Fellay R, Talmont F, Kamalaprija P, Burger U, RelićB,PromeJC,BroughtonWJ<br />

(1995) Involvement of nodS in N-methylation and nodU in 6-O-carbamoylation of<br />

Rhizobium sp. NGR234 Nod factors. J Biol Chem 270:22968–22973<br />

Kaneko T, Nakamura Y, Sato S, Asamizu E, Kato T, Sasamoto S, Watanabe A, Idesawa K,<br />

Ishikawa A, Kawashima K, Kimura T, Kishida Y, Kiyokawa C, Kohara M, Matsumoto M,<br />

Matsuno A, Mochizuki Y, Nakayama S, Nakazaki N, Shimpo S, Sugimoto M, Takeuchi C,<br />

Yamada M, Tabata S (2000) Complete genome structure of the nitrogen-fixing symbiotic<br />

bacterium Mesorhizobium loti. DNA Res 7:331–338<br />

Kaneko T, Nakamura Y, Sato S, Minamisawa K, Uchiumi T, Sasamoto S, Watanabe A,<br />

IdesawaK,IriguchiM,KawashimaK,KoharaM,MatsumotoM,ShimpoS,TsuruokaH,<br />

Wada T, Yamada M, Tabata S (2002) Complete genomic sequence of nitrogen-fixing<br />

symbiotic bacterium Bradyrhizobium japonicum USDA110. DNA Res 9:189–197<br />

Kobayashi H, Naciri-Graven Y, Broughton W, Perret X (2004) Flavonoids induce temporal<br />

shifts in gene expression of nod-box controlled loci in Rhizobium sp. NGR234. Mol<br />

Microbiol 51:335–347<br />

Krause A, Doerfel A, Göttfert M (2002) Mutational and transcriptional analysis of the<br />

type III secretion system of Bradyrhizobium japonicum. Mol Plant Microbe Interact<br />

15:1228–1235<br />

Krishnan HB (2002) NolX of Sinorhizobium fredii USDA257, a type III-secreted protein<br />

involved in host range determination, Is localized in the infection threads of cowpea<br />

(Vigna unguiculata [L.] Walp) and soybean (Glycine max [L.] Merr.) nodules. J Bacteriol<br />

184:831–839<br />

Krishnan HB, Pueppke SG (1993) Flavonoid inducers of nodulation genes stimulate Rhizobium<br />

fredii USDA257 to export proteins into the environment. Mol Plant Microbe<br />

Interact 6:107–113<br />

Krishnan HB, Kuo C-I, Pueppke SG (1995) Elaboration of flavonoid-induced proteins by<br />

the nitrogen-fixing soybean symbiont Rhizobium friedii is regulated by both nodD1 and


86 M.M. Saad et al.<br />

nodD2, and is dependent on the cultivar-specificity locus, nolXWBTUV. Microbiology<br />

141:2245–2251<br />

Krishnan HB, Lorio J, Kim WS, Jiang G, Kim KY, DeBoer M, Pueppke SG (2003) Extracellular<br />

proteins involved in soybean cultivar-specific nodulation are associated with pilus-like<br />

surface appendages and exported by a type III protein secretion system in Sinorhizobium<br />

fredii USDA257. Mol Plant Microbe Interact 16:617–625<br />

Marie C, Broughton WJ, Deakin WJ (2001) Rhizobium type III secretion systems: legume<br />

charmers or alarmers? Curr Opin Plant Biol 4:336–342<br />

Marie C, Deakin WJ, Viprey V, Kopcinska J, Golinowski W, Krishnan HB, Perret X,<br />

Broughton WJ (2003) Characterization of Nops, nodulation outer proteins, secreted<br />

via the type III secretion system of NGR234. Mol Plant Microbe Interact 16:743–751<br />

Moreira LM, Becker JD, Puhler A, Becker A (2000) The Sinorhizobium meliloti ExpE1<br />

protein secreted by a type I secretion system involving ExpD1 and ExpD2 is required<br />

for biosynthesis or secretion of the exopolysaccharide galactoglucan. Microbiology<br />

146:2237–2248<br />

Peabody CR, Chung YJ, Yen MR, Vidal-Ingigliardi D, Pugsley AP, Saier MH (2003) Type<br />

II protein secretion and its relationship to bacterial type IV pili and archaeal flagella.<br />

Microbiology 149:3051–3072<br />

Perret X, Freiberg C, Rosenthal A, Broughton WJ, Fellay R (1999) High-resolution transcriptional<br />

analysis of the symbiotic plasmid of Rhizobium sp. NGR234. Mol Microbiol<br />

32:415–425<br />

Perret X, Staehelin C, Broughton WJ (2000) Molecular basis of symbiotic promiscuity.<br />

Microbiol Mol Biol Rev 64:180–201<br />

Pueppke SG, Broughton WJ (1999) Rhizobium sp. strain NGR234 and R. fredii USDA257<br />

share exceptionally broad, nested host ranges. Mol Plant Microbe Interact 12:293–318<br />

Pugsley AP (1993) The complete general secretory pathway in Gram-negative bacteria.<br />

Microbiol Rev 57:50–108<br />

Relić B, Talmont F, Kopcinska J, Golinowski W, Prome JC, Broughton WJ. (1993) Biological<br />

activity of Rhizobium sp. NGR234 Nod-factors on Macroptilium atropurpureum. Mol<br />

Plant Microbe Interact 6:764–774<br />

Relić B, Perret X, Estrada-Garcia MT, Kopcinska J, Golinowski W, Krishnan HB, Pueppke SG,<br />

Broughton WJ (1994) Nod factors of Rhizobium are a key to the legume door. Mol<br />

Microbiol 13:171–178<br />

Rossier O, Van den Ackerveken G, Bonas U (2000) HrpB2 and HrpF from Xanthomonas are<br />

type III-secreted proteins and essential for pathogenicity and recognition by the host<br />

plant. Mol Microbiol 38:828–838<br />

Sandkvist M (2001) Biology of type II secretion. Mol Microbiol 40:271–283<br />

Spaink HP, Sheeley DM, van Brussel AA, Glushka J, York WS, Tak T, Geiger O, Kennedy EP,<br />

Reinhold VN, Lugtenberg BJ (1991) A novel highly unsaturated fatty acid moiety<br />

of lipo-oligosaccharide signals determines host specificity of Rhizobium Nature<br />

354:125–130<br />

Stacey G, So JS, Roth LE, Bhagya Lakshmi SK, Carlson RW (1991) A lipopolysaccharide mutant<br />

of Bradyrhizobium japonicum that uncouples plant from bacterial differentiation.<br />

Mol Plant Microbe Interact 4:332–340<br />

Streit WR, Schmitz RA, Perret X, Staehelin C, Deakin WJ, Raasch C, Liesegang H,<br />

Broughton WJ (2004) An evolutionary hot spot: the pNGR234b replicon of Rhizobium<br />

sp. Strain NGR234. J Bacteriol 186:535–542<br />

Sullivan JT, Trzebiatowski JR, Cruickshank RW, Gouzy J, Brown ST, Elliot RM, Fleetwood DJ,<br />

McCallum NG, Rossbach U, Stuart GS, Weaver JE, Webby RJ, de Bruijn FJ, Ronson cw<br />

(2002) Comparative sequence analysis of the symbiosis island of Mesorhizobium loti<br />

StrainR7A. J Bacteriol 184:3086–3095


5 Protein secretion by Rhizobia 87<br />

Sutton JM, Lea EJ, Downie JA (1994) The nodulation-signalling protein NodO from Rhizobium<br />

leguminosarum biovar viciae forms ion channels in membranes. Proc Natl Acad<br />

Sci USA 91:9990–9994<br />

Sutton JM, Peart J, Dean G, Downie JA (1996) Analysis of the C-terminal secretion signal<br />

of the Rhizobium leguminosarum nodulation protein NodO; a potential system for the<br />

secretion of heterologous proteins during nodule invasion. Mol Plant Microbe Interact<br />

9:671–680<br />

Thanassi DG, Hultgren SJ (2000) Multiple pathways allow protein secretion across the<br />

bacterial outer membrane. Curr Opin Cell Biol 12:420–430<br />

Trinick MJ (1980) Relationships amongst the fast-growing rhizobia of Lablab purpureus,<br />

Leucaena leucocephala, Mimosa spp., Acacia farnesiana and Sesbania grandiflora and<br />

their affinities with other rhizobial groups. J Appl Bacteriol 49:39–53<br />

Van Rhijn P, Luyten E, Vlassak K, Vanderleyden J (1996) Isolation and characterization<br />

of a pSym locus of Rhizobium sp. BR816 that extends nodulation ability of narrow<br />

host range Phaseolus vulgaris symbionts to Leucaena leucocephala Mol Plant Microbe<br />

Interact 9:74–77<br />

Van Spronsen PC, Bakhuizen R, van Brussel AA, Kijne JW (1994) Cell wall degradation during<br />

infection thread formation by the root nodule bacterium Rhizobium leguminosarum<br />

is a two-step process. Eur J Cell Biol 64:88–94<br />

Viprey V, Del Greco A, Golinowski W, Broughton WJ, Perret X (1998) Symbiotic implications<br />

of type III protein secretion machinery in Rhizobium. Mol Microbiol 28:1381–1389<br />

Vlassak KM, de Wilde P, Snoeck C, Luyten E, van Rhijn P, Vanderleyden J (1998) The<br />

Rhizobium sp. BR816 nodD3 gene is regulated by a transcriptional regulator of the<br />

AraC/XylS family. Mol Gen Genet 258:558–561<br />

Walker SA, Downie JA (2000) Entry of Rhizobium leguminosarum bv. viciae into root hairs<br />

requires minimal Nod factor specificity, but subsequent infection thread growth requires<br />

nodO or nodE Mol Plant Microbe Interact 13:754–762<br />

Winans SC, Burns DL, Christie PJ (1996) Adaptation of a conjugal transfer system for the<br />

export of pathogenic macromolecules. Trends Microbiol 4:64–68<br />

Wolfgang M, van Putten JP, Hayes SF, Dorward D, Koomey M (2000) Components and<br />

dynamics of fibres formation define a ubiquitous biogenesis pathway for bacterial pili.<br />

EMBO J 19:6408–6418


6<br />

Research on Endophytic Bacteria:<br />

Recent Advances with Forest Trees<br />

Richa Anand, Leslie Paul, Chris Chanway<br />

6.1<br />

Introduction<br />

Plantscanbeconsideredascomplex microecosystemsthatprovidedifferent<br />

habitats to a variety of microorganisms. These habitats are represented<br />

by the plant external surfaces as well as internal tissues (McInroy and<br />

Kloepper 1994). Whereas the importance of microbial colonisation of plant<br />

surfaces in plant growth promotion has been well understood for a long<br />

time, interior tissue colonisation was, until recently, largely perceived as<br />

being related only to the perpetuation of systemic diseases. It is now well<br />

known that tissues of healthy plants are also colonised internally by various<br />

microorganisms that establish neutral or, more interestingly, beneficial<br />

interactions with their host plants. The term “endophyte” is commonly<br />

used to describe such microorganisms.<br />

Although a variety of definitions have been applied to the term “endophyte”,<br />

it refers mainly to bacteria and fungi that live inside plant tissues<br />

without causing disease (Wilson 1995; see Chap. 1 by Schulz and Boyle).<br />

Whether or not latent pathogens can be considered endophytes has been<br />

a major topic of debate in the general acceptance of this definition (Misaghi<br />

and Donndelinger 1990; James and Olivares 1997; see Chap. 1 by Schulz<br />

and Boyle).<br />

The best-characterised microbial endophytes are fungi of the Balansiaceae,<br />

for which the most compelling evidence of plant–microbe mutualism<br />

has been provided (Clay 1988; Schardl et al. 2004). Some of the<br />

non-balansiaceous endophytic fungi are also mutualistic with their hosts<br />

(Carroll 1988; see Chap. 15 by Schulz), and produce compounds that render<br />

plant tissues less attractive to herbivores, while other strains may increase<br />

Chris Chanway: Faculty of Land and Food Systems, Faculty of Forestry, University<br />

of British Columbia, Vancouver, British Columbia, Canada V6T 1Z4, E-mail:<br />

cchanway@interchg.ubc.ca<br />

Richa Anand, Leslie Paul: Faculty of Land and Food Sciences, Systems, University of British<br />

Columbia, Vancouver, British Columbia, Canada V6T 1Z4<br />

Leslie Paul: Department of Forest Mycology and Pathology, Swedish University of Agricultural<br />

Sciences (SLU), Box 7026, 750 07 Uppsala, Sweden (Current Address)<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


90 R. Anand et al.<br />

host plant drought resistance. In return, fungal endophytes are thought to<br />

benefit from the comparatively nutrient rich, buffered environment inside<br />

plants.<br />

Apart from fungi, bacteria belonging to various genera have also been<br />

shown to exist inside plants without causing apparent disease symptoms.<br />

Some of these bacteria are known to impart benefits to their host plants<br />

by the same mechanisms as their soil- or rhizosphere-colonising counterparts.<br />

The primary mechanisms thought to lead to beneficial effects for<br />

the plant are nitrogen fixation (Boddey and Döbereiner 1995) and biocontrol<br />

of pathogenic and detrimental microorganisms, either through<br />

direct antagonism of pathogens or by inducing systemic resistance to such<br />

organisms (Hallman et al. 1997). Other known mechanisms by which beneficial<br />

bacteria can have a positive influence on plant performance are the<br />

production or stimulation of plant growth hormones and facilitation of<br />

nutrient uptake [see Chaps. 3 (Kloepper and Ryu) and 4 (Berg and Hallmann)].<br />

Since their first reported isolation from potato plants (Tervet and Hollis<br />

1948; Hollis 1951), all the information available on endophytic bacteria has<br />

been derived from studies on plant species of agricultural and horticultural<br />

importance. The endophytic bacteria of rice (Reinhold-Hurek and Hurek<br />

1998), corn (Triplett 1996) and sugarcane (Döbereiner et al. 1995) are by<br />

far the best studied so far. In contrast to these crop species, very much<br />

less is known about bacterial endophytes of trees. Some trees survive and<br />

grow well in very difficult terrain under extreme conditions, for example<br />

lodgepole pine (Pinus contorta Dougl. var. latifolia)indryinteriorregions<br />

of British Columbia and western Alberta, Canada, as well as Tecomella<br />

undulata (Bignoniaceae) in the extremely arid deserts of northwestern<br />

India. It is possible that endophytic bacteria that enhance host survival and<br />

growth in exchange for protection in the relatively buffered environment of<br />

internal plant tissues may be involved under such extreme environmental<br />

conditions (Law and Lewis 1983).<br />

Although the realisation of this possibility has led to occasional reports<br />

of endophytic bacteria in asymptomatic angiosperm and gymnosperm tree<br />

species, little is known about their diversity and influence on plant growth.<br />

The earliest report of bacterial endophytes from trees was from Gardner<br />

et al. (1982), who isolated representatives of 13 genera from xylem fluid of<br />

rough lemon rootstock, and found population sizes ranging from 10 2 –10 4<br />

colony forming units (cfu) g −1 xylem fluid. Only 48 of the 850 isolates<br />

turned out to be phytopathogenic, but the role of the other 802 isolates was<br />

not determined. Similarly, several strains of Pseudomonas syringae were<br />

isolated and characterized from inside pear seedlings by Whitesides and<br />

Spotts (1991), but their exact role could not be determined.


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 91<br />

The procedures of isolation and identification of endophytic bacteria<br />

fromtreesisthesameastheirisolationfromagronomiccrops,andsuffers<br />

from the same limitations observed in crop species, e.g. the difficulty, or<br />

perhaps impossibility, of absolute surface sterilisation of external plant<br />

tissues as well as our inability to culture many bacteria we know to exist.<br />

The impact of these problems can be reduced by the use of standardised<br />

protocols and molecular techniques (James 2000; see Chap. 17 by Hallmann<br />

et al.). The major difficulty, therefore, lies in the evaluation of the effects of<br />

these bacteria on their host trees, owing to the long tree life-cycle and a lack<br />

of detailed physiological information on trees, particularly forest trees.<br />

The following sections provide a detailed review of our current knowledge<br />

about bacterial endophytes of forest trees, the mechanisms by which<br />

they benefit their host plants, and their potential application in the practice<br />

of sustainable forestry.<br />

6.2<br />

Bacterial Endophytes of Forest Trees<br />

Although limited, the results of research on endophytic bacteria and their<br />

role in growth promotion of forest trees so far are very encouraging and<br />

will hopefully draw more attention to this developing area of study. Brooks<br />

et al. (1994) conducted an extensive study in which endophytic bacteria<br />

were isolated from surviving live oak (Quercus fusiformis)inTexas,where<br />

oak wilt is epidemic, and evaluated as potential biological control agents<br />

for the disease. Of the 889 bacterial isolates tested, 183 showed in vitro inhibition<br />

of the pathogen Ceratocystis fagacearum. Six isolates were further<br />

evaluated for colonisation of containerised Spanish oak (Quercus texana)<br />

and live oak. Interestingly, in containerised live oaks inoculated with the<br />

oak wilt pathogen, preinoculation with 15 isolates of Pseudomonas denitrificans<br />

reduced the number of diseased trees by 50% and decreased the<br />

percentage of crown loss by 17%. In a subsequent trial, no reduction in<br />

numbersofdiseasedtreeswasobservedbutpreinoculationwiththesame<br />

isolates of P. denitrificans or a strain of Pseudomonas putida significantly<br />

reduced crown loss. These results clearly established the potential of such<br />

endophytic bacteria as pre-plantation nursery treatments for wilt control.<br />

Several endophytic aerobic heterotrophic bacteria belonging to the genera<br />

Bacillus, Curtobacterium, Pseudomonas, Stenotrophomonas, Sphingomonas,<br />

Enterobacter, andStaphylococcus, havealsobeenisolatedfrom<br />

phloem tissue of roots and branches of elm trees (Ulmus spp.: Mocali<br />

et al. 2003). An attempt was also made to determine the correlation between<br />

the seasonal fluctuations in the structure of the endophytic bacterial


92 R. Anand et al.<br />

community and phytoplasma disease infection of these trees; however, no<br />

consequential effect of the bacterial community on phytoplasmosis of elm<br />

trees could be demonstrated (Mengoni et al. 2003).<br />

Apart from these studies on bacterial endophytes of deciduous trees,<br />

most other reports of endophytic bacteria and their role in tree growth promotion<br />

are from studies on various conifer tree species conducted mostly<br />

by our research group. Our interest in endophytic bacteria of conifers has<br />

been largely inspired by the immense commercial, social, and environmental<br />

importance of forestry in Canada and the rest of North America and the<br />

fact that conifers are the most dominant trees in the temperate forests of<br />

this region.<br />

6.3<br />

Endophytic Bacteria of Conifers<br />

Conifers are members of the plant division Coniferophyta (2 Domain classification),<br />

which are characterised by naked seeds borne in specialised<br />

sporophylls or cones. Their vascular tissues differ from angiosperms in<br />

not having xylem vessels, and companion cells in phloem. The division is<br />

comprisedof550speciesspreadoversevenfamilies,eachdatingbacktothe<br />

Mesozoic era. Distributed throughout the world with extensive latitudinal<br />

and longitudinal ranges, conifers are of great commercial and ecological<br />

value.<br />

Traditionally, fungi, particularly mycorrhizae, were considered to be the<br />

only microorganisms that could exert a positive influence on the growth<br />

and survival of forest trees. The continuity of this trend until now is evident<br />

from the results of keyword searches for “endophytic bacteria + conifers”<br />

in all well known scientific databases.<br />

Although some confirmed reports of conifer tree growth promotion by<br />

naturally occurring soil and rhizosphere bacteria were available (Pokojska-<br />

Burdziej 1982; Chanway and Holl 1992, 1993, 1994; O’Neill et al. 1992), the<br />

mechanisms employed by these bacteria for growth promotion could not be<br />

determined. It was generally believed that the primary mechanism of plant<br />

growth promotion by these bacteria was only indirect, i.e. by facilitating<br />

the establishment and growth of mycorrhizae (Fitter and Garbaye 1994).<br />

The focus of research on endophytic microflora of conifers thus remained<br />

on fungi, even after the importance of endophytic bacteria had been well<br />

established in agronomic crop species.<br />

In an initial study of conifer root-associated bacteria, O’Neill et al. (1992)<br />

isolated 22 strains from surface-sterilised roots of naturally regenerating<br />

white x Engelmann (Picea glauca x P. engelmannii) hybrid spruce seedlings.<br />

A range of effects on seedling growth in a greenhouse-screening assay


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 93<br />

using spruce were found: three strains were inhibitory, five strains were<br />

stimulatory and the remaining strains had no significant effect on seedling<br />

growth (O’Neill et al. 1992). Based on the magnitude and consistency of<br />

seedling growth effects, the two best plant growth-promoting endophytes<br />

were identified and selected for further study: one isolate was Pseudomonas<br />

putida and the other belonged to Staphylococcus. While the positive effect<br />

of both of these strains on plant growth was reproducible in the greenhouse,<br />

a field trial with two ecotypes of 1-year-old spruce seedlings planted at three<br />

different reforestation sites yielded mixed results (Chanway and Holl 1993).<br />

For example, P. putida enhanced seedling growth of only one of two spruce<br />

ecotypes planted at two of three reforestation sites. In addition, it had<br />

inhibitory effects in three of the spruce ecotype x planting site treatment<br />

combinations.<br />

Evaluation of gymnosperm bacterial endophytes was only a small part<br />

of a larger project designed to characterise gymnosperm root-associated<br />

bacterial (i.e. external and internal) colonists (O’Neill et al. 1992; Chanway<br />

and Holl 1992, 1994). Therefore, our group undertook a subsequent bacterial<br />

isolation and screening program emphasising endophytic bacteria<br />

as possible tree seedling growth-promoting agents (Chanway et al. 1994,<br />

1997). As seen in our earlier work (O’Neill et al. 1992), several bacterial<br />

strains isolated from surface-sterilised roots of white x Engelmann hybrid<br />

spruce seedlings caused reproducible spruce seedling biomass increases of<br />

up to 36% 2 months after seed was sown and inoculated in greenhouse<br />

trials (Chanway et al. 1994). Three of these strains belonged to Paenibacillus,<br />

three were actinomycetes, most likely Streptomyces spp., and one was<br />

a Phyllobacterium. An additional strain that performed well in greenhouse<br />

assays could not be identified with certainty.<br />

In addition, the seedling growth promotion efficacy of some of these<br />

strains was altered significantly when assays were conducted in the presence<br />

of a small amount (2% v/v) of forest soil known to contain seedling growth<br />

inhibiting organisms (i.e. minor pathogens). One of the endophytic actinomycetes<br />

(isolate W2) as well as the Phyllobacterium isolate (W3) clearly<br />

stimulated spruce seedling growth only in the absence of forest soil. In its<br />

presence, seedling growth was inhibited, as it was when forest soil alone<br />

was used. These results suggested that growth promotion by W2 and W3<br />

occurred via a mechanism unrelated to biocontrol of minor pathogens, and<br />

mayhaveinvolvedoneofthedirectplantgrowthpromotionmechanisms,<br />

possibly production of plant growth regulators (Kloepper 1993; Glick 1995;<br />

Chanway 1997). Interestingly, actinomycete isolate N1 and Bacillus isolate<br />

N4 stimulated seedling growth only in the presence of forest soil, which<br />

suggested that these strains acted through a biocontrol mechanism, either<br />

by direct antagonism or by inducing systemic resistance in the host plant.<br />

Elucidation of these possibilities requires further experimentation.


94 R. Anand et al.<br />

We have also looked for bacterial endophytes in lodgepole pine. After isolation<br />

of several bacterial strains and screening trials for effects on seedling<br />

growth, we identified a plant-growth-promoting Bacillus polymyxa (now<br />

Paenibacillus, Ash et al. 1993) strain (Pw2) that originated from internal<br />

root tissues of a naturally regenerating 2- to 3-year-old pine seedling<br />

(Shishido et al. 1995). Our studies indicate that Pw2 can colonise external<br />

and internal pine and spruce root tissues after seed or root inoculation.<br />

Colonisation of internal root tissues may depend on lateral root development,<br />

and results in endophytic bacterial population sizes approaching<br />

10 6 cfu g −1 fresh root tissue (Shishido et al. 1995; Chanway 1997; Shishido<br />

1997). In addition, using a surface-sterilisation, dilution plating assay as<br />

well as immunofluorescence microscopy, a rifamycin-resistant derivative<br />

of this strain, Pw2-R, was shown to be capable of colonising internal pine<br />

and white x Engelmann hybrid spruce stem tissues after soil or root inoculation<br />

(Chanway et al. 2000). Five months after root inoculation, internal<br />

stem bacterial populations reached 10 5 cfu g −1 fresh stem tissue (Shishido<br />

1997).<br />

In order to examine the effects of endophytes on conifer plant growth<br />

and to investigate the host specificity of bacterial endophytes in terms of<br />

the ability to promote growth of inoculated host plants other than the ones<br />

from which they were initially isolated, initial field trials with P. polymyxa<br />

strain Pw2-R and Pseudomonas chloroaphis strain Sm3-RN, another bacterial<br />

endophyte capable of stimulating white x Engelmann hybrid seedling<br />

growth in the greenhouse (Chanway et al. 1997), were also performed.<br />

Two years after bacterial inoculation and planting at nine sites in British<br />

Columbia and Alberta, Canada, white x Engelmann hybrid spruce treated<br />

with strain Pw2-R (initially isolated from pine) showed mean biomass increases<br />

up to 33% above controls at seven of the nine sites, but increases<br />

were significant at only one site. In contrast, Pseudomonas strain Sm3-RN<br />

(isolated from white x Engelmann hybrid spruce) caused significant white<br />

x Engelmann hybrid spruce biomass increases of up to 57% at three of<br />

the nine sites but a significant decrease in spruce biomass at one site. Site<br />

productivity was not correlated with plant growth promotion or inhibition.<br />

Population sizes of Pw2-R and Sm3-RN were generally below the assay<br />

detection limit of ca. 10 2 cfu g −1 plant tissue, which led us to question how<br />

effectively internal plant tissues were colonised at the onset of the experiment.<br />

Therefore, we also evaluated seedlings that were inoculated with<br />

strainsPw2-RandSm3-RNandgrowninthegreenhousefor4months<br />

before planting at four of the reforestation sites described above (Shishido<br />

and Chanway 2000). The period of growth in the greenhouse facilitated<br />

internal tissue colonisation by these microorganisms so that mean internal<br />

root populations reached ca. 10 3 –10 4 cfu g −1 tissue in the greenhouse. As


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 95<br />

expected, mean seedling biomass also increased due to bacterial inoculation<br />

in the greenhouse. Because seedling growth responses in the field<br />

would be inseparable from those that occurred in the greenhouse, simple<br />

measurement of biomass accumulation after a period of growth in the<br />

field would yield spurious results. We evaluated plant growth using relative<br />

growth rates (RGRs), in which plant growth increments over time are expressed<br />

as a proportion of the biomass that existed at some previous time<br />

in the plant’s life (Hunt 1982).<br />

In general, after the first growing season, RGRs of seedlings containing<br />

endophytic bacteria were greater than those of control seedlings at all four<br />

planting sites (Shishido and Chanway 2000). In some cases, RGRs of inoculated<br />

plants were double the control value. This was particularly interesting<br />

inviewofresultswithseedlingsthatweinoculatedandplantedimmediately<br />

at the same sites. At two of the four sites, seedlings inoculated at the time<br />

of planting (i.e., with no greenhouse growth period) did not respond to<br />

bacterial treatment, and in one case responded negatively. However, shoot<br />

androotRGRsofseedlingspretreatedinthegreenhousebeforeplantingat<br />

the same sites were 23–132% greater than controls, and endophytic populations<br />

in root tissues of between 10 2 and 4×10 4 cfu g −1 plant tissue were<br />

detected in seedlings at three of the four sites. Similar effects on establishment<br />

and functioning of bacterial endophytes were observed by Brooks<br />

et al. (1994) in wilt-infested oak trees. These results suggest that a period<br />

of growth under a controlled environment to facilitate establishment of<br />

endophytic bacterial populations may be an important step in successful<br />

application of plant-growth-promoting bacterial endophytes in forestry.<br />

It has also been demonstrated that the benefits of pre-outplanting inoculation<br />

of seedlings with bacterial endophytes can be maximised by careful<br />

matching of the inoculant bacterial strain with outplanting sites (Chanway<br />

et al. 2000). However, much research into site quality and plant growth<br />

responseswillberequiredbeforereliablerecommendationscanbemade.<br />

In addition, much more research is warranted to answer the many questions<br />

regarding the entry and operation of endophytic bacteria in conifers,<br />

some of which are being actively pursued by our group.<br />

6.4<br />

ModesandSitesofEntry<br />

Endophytic bacteria have been shown to be able to gain entry in plants<br />

through wounded as well as intact tissues (Sprent and James 1995; see<br />

Chap. 18 by Bloemberg et al.). In an attempt to understand the modes and<br />

sites of entry of endophytic bacteria, Timmusk and Wagner (1999) followed<br />

the colonisation of a green fluorescent protein (gfp)-tagged endophytic


96 R. Anand et al.<br />

strain of Paenibacillus polymyxa in Arabidopsis thaliana; theyobserved<br />

aslightdegradationoftheroottipswithin5hofinoculation.Itwasfound<br />

that P. polymyxa hadtwopreferredzonesofinfection.Thefirstislocated<br />

at the root tip in the zone of elongation, which sometimes results in the<br />

loss of the root cap. The other colonisation region was observed in the<br />

differentiation zone. Similar colonisation zones have been reported for<br />

other endophytes, e.g. Azoarcus by Hurek et al. (1994), who suggested<br />

that plant cells were destroyed after bacteria had penetrated cell walls.<br />

Perhaps this is the reason why most endophytic bacteria are limited to<br />

the intercellular spaces inside tissues. However it is not clear how they are<br />

stopped from entering cells and causing necrosis.<br />

To determine which microbial characteristic(s) facilitate entrance of<br />

bacterial endophytes into plant tissues, we compared the biochemical capabilities<br />

of the endophytic Paenibacillus polymyxa strain Pw2 with those<br />

of another plant-growth promoting, non-endophytic strain, P. polymyxa<br />

L6-16R. Interestingly, strain L6-16R is unable to enter plant tissues even<br />

when co-inoculated with an endophytic microorganism (Shishido et al.<br />

1995; Bent and Chanway 1997). According to Biolog analysis, both strains<br />

possessed similar metabolic capabilities with some potentially important<br />

exceptions (Shishido et al. 1995). For example, strain Pw2-R was able to<br />

metabolise sorbitol, but strain L6-16R was not. Mavingui et al. (1992) found<br />

that, in general, P. polymyxa strains isolated from the rhizoplane of wheat<br />

(Triticum aestivum L.) were capable of metabolising sorbitol while rhizosphere<br />

and non-rhizosphere isolates were not. They hypothesised that<br />

intense competition for oxygen would occur on the root surface due to root<br />

respiration, which would result in selection pressure for bacteria capable<br />

of anaerobic growth on highly reduced, scarce substrates, such as sorbitol.<br />

In addition, strain Pw2 was able to metabolise d-melezitose, a sugar that<br />

has been detected in the sap of conifers (Lehninger 1975). However, the<br />

occurrence of sorbitol and d-melezitose in lodgepole pine root tissues and<br />

their utilisation by other Paenibacillus root endophytes must be demonstrated<br />

before a role for these substrates in internal root colonisation by<br />

Paenibacillus can be postulated with greater confidence.<br />

To facilitate root colonisation, it is logical to suspect that root endophytic<br />

bacteria may also possess the ability to metabolise structural components of<br />

plant cells. In particular, the ability to metabolise pectin (polygalacturonic<br />

acids), a major component of the middle lamellae of plant cell walls, has<br />

been proposed to at least partly explain why bacterial root endophytes<br />

are often found in the root cortex intercellularly (Balandreau and Knowles<br />

1978; Baldani and Döbereiner 1980). Both strains L6 and Pw2 possessed<br />

pectolytic activity in vitro, but only strain Pw2 was able to metabolise dgalacturonic<br />

acid (Shishido et al. 1995), the primary monomeric component<br />

of pectin (Paul and Clark 1989).


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 97<br />

It is not clear whether the capability of strain Pw2 to metabolise monomeric<br />

galacturonic acid after breakdown of the pectin polymer was related<br />

to its ability to enter root tissues. However, breakdown products of plant<br />

cell walls are known to induce systemic disease responses in plants (Brock<br />

et al. 1994), which leads to the possibility that Pw2 avoids plant defence<br />

mechanisms by metabolising cell wall components before they elicit a defence<br />

response by the host plant. This possibility also requires further<br />

investigation.<br />

If, in fact, the entry of Pw2 in plant roots is facilitated by its capability to<br />

metabolise the primary components of the cell wall, the question as to why<br />

it does not cause necrosis of interior tissues, also needs to be answered.<br />

6.5<br />

Mechanisms of Plant Growth Promotion<br />

Unlike symbiotic rhizobia, mechanisms of plant growth promotion by<br />

plant growth-promoting rhizobacteria (PGPR) vary greatly, and have been<br />

broadly categorised into two groups, direct and indirect (Kloepper et al.<br />

1989; see Chap. 3 by Kloepper and Ryu). Direct plant growth mechanisms<br />

may involve nitrogen fixation (Cavalcante and Döbereiner 1988), production<br />

of plant growth regulators and antibiotics, or increased availability of<br />

plant growth-limiting nutrients. Indirect mechanisms may involve suppression<br />

of deleterious microorganisms as well as enhancement of mutualisms<br />

between host plants and other symbionts such as mycorrhizae (Kloepper et<br />

al. 1989). Similar to other aspects of studies on endophytic bacteria, there is<br />

a great deal of information on the mechanisms of plant growth promotion<br />

employed by these bacteria in agronomic crops (Lodewyckx et al. 2002).<br />

In the case of conifers, it was generally believed that these plants could derive<br />

benefits from bacteria only indirectly through their mycorrhizal symbionts<br />

(Fitter and Garbaye 1994). However, growth studies on lodgepole<br />

pine seedlings (Chanway and Holl 1991; Shishido et al. 1996) and hybrid<br />

spruce (Picea glauca x P. engelmannii) (Shishido et al. 1996) co-inoculated<br />

with PGPR and mycorrhizal fungi have clearly shown that growth promotionoftheseconifersbyPGPRisindependentofthemycorrhizalstatusof<br />

the seedlings.<br />

Despite many efforts, determination of the exact mechanisms of conifer<br />

growth promotion by PGPR has not been possible. Paenibacillus polymyxa<br />

strain L6-16R was shown to produce cytokinins (Holl et al. 1988), and this<br />

propertywasadvancedasalikelyexplanationofpinegrowthpromotion<br />

mediated by this strain.<br />

A detailed study was also conducted to determine the mechanisms of<br />

growth promotion of spruce by six Paenibacillus and Pseudomonas strains,


98 R. Anand et al.<br />

including the endophyte, B. polymyxa Pw2 (Shishido 1999). It could only<br />

be concluded that more than one mechanism was responsible for growth<br />

promotion. Production of plant growth promoters and enhancement of<br />

nutrient uptake were designated as the most likely of these mechanisms.<br />

Strain Pw2 isolated from lodgepole pine (Shishido 1996) possessed diazotrophic<br />

properties. This led to the intriguing possibility that lodgepole<br />

pine harbours a systemically endophytic nitrogen-fixing bacterial population,<br />

similar to that found in sugar cane (Boddey et al. 1995), which would<br />

explain its ability to grow, and even thrive, under nitrogen-deficient conditions<br />

in the absence of significant rhizospheric nitrogen fixation (Binkley<br />

1995). Indeed, the 15 N/ 14 N ratio of pine foliage in a central coast forest<br />

in British Columbia devoid of nitrogen-fixing species was observed to be<br />

low enough to suggest that biological nitrogen fixation supplies plant N<br />

(F.B. Holl, personal communication).<br />

However, nitrogen fixation could not be shown to be the primary mechanism<br />

of growth promotion by P. polymyxa strain Pw2-R, since seedlings inoculated<br />

with it failed to support sufficient rhizosphere acetylene reduction<br />

activity (ARA) even after 48 h of incubation with acetylene (Shishido 1997).<br />

Interestingly, similar limitations were encountered by Rhodes-Roberts<br />

(1981) and Achouak et al. (1999) while working with other strains of<br />

P. polymyxa. However, they were able to measure the nitrogen gains of<br />

seedlings by microkjeldahl analysis, which led them to suggest that the<br />

acetylene reduction assay is not always able to provide positive results for<br />

the nitrogen-fixing ability of P. polymyxa. Therefore, conclusions on the<br />

occurrence of N2 fixationinvivoshouldbedrawnfromanumberoflines<br />

of evidence, including a positive nitrogenase activity test (acetylene reduction<br />

assay), 15 N dilution and detection of conserved nif genes in the<br />

purported diazotrophic endophyte.<br />

We have also observed nitrogen-fixing bacteria inside what can only be<br />

describedasauniqueandenigmatictypeofmycorrhizaeonlodgepolepine,<br />

first described by Zak (1971) on Douglas-fir (Pseudotsuga menziesii) roots.<br />

These mycorrhizal structures, often referred to as tuberculate mycorrhizae,<br />

look more like leguminous root nodules than mycorrhizae (Fig. 6.1). They<br />

are fully enclosed subterranean “nodules” or tubercles attached to the tree<br />

root system, with hundreds more typical mycorrhizal root tips crowded<br />

inside the outer covering, or peridium (Fig. 6.2). Nitrogen-fixing bacteria<br />

have been previously detected on the peridium (Li et al. 1992), but more<br />

recently, in our laboratory, a limited number of strains representing four<br />

diazotrophic bacterial species have been detected inside the peridium,<br />

colonising the fungal hyphae within the tubercle (unpublished data). It is<br />

yet to be demonstrated that these endophytic diazotrophs fix N2 in situ, let<br />

alone transfer it to the host plant, but these intriguing possibilities remain<br />

to be evaluated.


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 99<br />

Fig.6.1. External morphology of tuberculate ectomycorrhizae on Pinus contorta roots. Bar<br />

5mm<br />

Fig.6.2. Cross section through a mature tubercle from P. contorta revealing mycorrhizal<br />

root tips (brown) and interstitial hyphae (arrow). Note pinnate radiated fan form and<br />

dichotomous branching of root tips within the tubercle. Bar 2mm<br />

Recently, we tested the hypothesis that diazotrophic endophytes isolated<br />

from internal tissues of immature and mature, naturally regenerated,<br />

lodgepole pine produce biologically significant amounts of N through


100 R. Anand et al.<br />

N-fixation under controlled environmental conditions. Entire lodgepole<br />

pine seedlings, as well as root, stem, and needle samples from trees were<br />

collected from 40- to 140-year-old stands near Williams Lake, British<br />

Columbia (52 ◦ 05 ′ N, 122 ◦ 54 ′ W, elevation 1,300 m) and Chilliwack Lake,<br />

British Columbia (49 ◦ 10 ′ N, 121 ◦ 57 ′ W, elevation 600 m). In addition, western<br />

red cedar (Thuja plicata Donn ex D. Don) samples were collected from<br />

a similar aged stand near Boston Bar, British Columbia (49 ◦ 50 ′ N, 121 ◦ 31 ′ W,<br />

elevation 600 m). Cedar samples were obtained in the same manner in<br />

which pine samples were collected except cedar stem samples from trees<br />

were obtained by cutting small wedges from stems using a pruning knife<br />

wiped down with 6% NaOCl prior to each sampling.<br />

Stem samples from mature trees were obtained by taking cores with<br />

a surface disinfested increment borer after shaving a thin layer of bark from<br />

the stem with a sterile scalpel. Root samples were surface-sterilised and<br />

triturated (Chanway et al. 2000) before endophytic bacteria were isolated<br />

by plating the resulting slurry.<br />

Four diazotrophic endophytes were isolated using this sampling procedure<br />

and 16S rDNA sequencing identified them as belonging to the genus<br />

Paenibacillus (strains P2b-2R, P18b-2R and C3b) as well as to the Flexibacter<br />

group (strain P19a-2R). The three strains with names beginning with<br />

“P”wereoriginallyisolatedfrompinetissuesfromtheWilliamsLakesite.<br />

Strain P2b was isolated from within the surface-sterilised stem of a pine<br />

seedling, strain P18b was isolated from within surface-sterilised needles<br />

of another pine seedling, and strain P19a was isolated from the internal<br />

stem tissue of a third pine seedling. Strain C3b was isolated from within the<br />

surface-sterilised stem of a western red cedar tree at the Boston Bar site.<br />

These microorganisms were then used to inoculate pine seed sown in<br />

glass tubes (150 mm × 25 mm in diameter) filled with a severely nitrogen<br />

deficient seedling growth medium to which a small amount (0.0576 g/l)<br />

Ca( 15 NO3)2 (5% 15 N label) was added to facilitate identification of foliar nitrogen<br />

originating from the growth medium versus the atmosphere. Other<br />

nutrients were added in amounts sufficient to support healthy plant growth.<br />

Bacterial inoculum was prepared by streaking frozen cultures onto plates<br />

of combined carbon medium (Rennie 1981). Following growth on plates,<br />

aloopfulofeachstrainwasseparatelyinoculatedintoitsown1lflask<br />

containing 500 ml CCM broth. Flasks were then secured on a rotary shaker<br />

(150 rpm; room temperature) and agitated for up to 2 days. All bacterial<br />

cultures were harvested by centrifugation (10,000g for 30 min), and resuspended<br />

in sterile phosphate buffer (SPB). Strains P2b-2R and C3b were<br />

resuspended to a density of ca. 10 7 cfu/ml and strains P18b-2R and P19a-2R<br />

were resuspended to a density of ca. 10 6 cfu/ml. For inoculation, 5.0 ml of<br />

each bacterial suspension was pipetted into separate tubes. Control seeds<br />

received 5.0 ml SPB. Tubes were placed in a growth chamber (Conviron


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 101<br />

CMP3244, Conviron Products Company, Winnipeg, MB). Photosyntheticallyactiveradiation(PAR)atcanopylevelwasca.300µmols<br />

−1 m −2 during<br />

an 18-h photoperiod, and 20 ◦ C/14 ◦ C day/night temperature cycle.<br />

Of the four diazotrophic endophytes, only inoculation with strain P2b-<br />

2R resulted in large, statistically significant increases of 27–66% in pine<br />

foliar nitrogen derived from the atmosphere in all three plant growth<br />

trials we have conducted to date (Table 6.1). Interestingly, strain P18b-<br />

2R, which was found to be phylogenetically very similar to P2b-2R, had no<br />

such effect on seedling foliar nitrogen. Clearly, minor genetic differences<br />

betweenendophyticstrainscanresultinprofoundlydifferenteffectson<br />

plant growth and foliar nitrogen derived from the atmosphere (NDFA).<br />

While large, statistically significant amounts of foliar NDFA were detected<br />

in all three trials with pine, there was no corresponding positive<br />

growth response in the first two experiments. Indeed, pine seedling growth<br />

was inhibited by strain P2b-2R compared to noninoculated controls in the<br />

first two trials (Bal 2003). That inoculated seedlings may derive nitrogen<br />

primarily from the atmosphere is an intriguing phenomenon; poor growth<br />

of inoculated seedlings may be an indication of the energetic cost of supporting<br />

nitrogen-fixing bacteria in the plant. Seedling growth reduction<br />

during the establishment of symbiosis is not uncommon, due to the energy<br />

diverted from the plant, and has been reported previously (Chanway<br />

and Holl 1991). This idea is supported by results from the third growth<br />

trial. By the end of the trial, control seedlings had a restricted growth rate,<br />

presumably due to N limitation, compared to the inoculated seedlings.<br />

P2b-2R-inoculated seedlings had greater biomass (47%) and total nitrogen<br />

(38%) compared to noninoculated controls, and derived 27% of foliar<br />

Nfromtheatmosphere.Intrials1and2,controlseedlingswereapparentlyscavengingthesmallamountofsoilNthatwasinitiallyprovided,<br />

and indeed outgrew the inoculated seedlings, which may have been in<br />

a “symbiosis development” mode. This tenet is hypothetical at this time<br />

and requires further investigation.<br />

Table 6.1. Percentages of nitrogen derived from the atmosphere (NDFA) in pine seedlings<br />

after inoculation with Paenibacillus polymyxa strain P2b-2R in three separate growth trials<br />

Growth Trial Duration (months after planting and inoculation) %NDFA a<br />

1 8 30<br />

2 9 66<br />

3 9 27<br />

aCalculated according to Rennie et al. (1978), where:<br />

�<br />

atom % 15N excess(inoculated plant)<br />

%NDFA = 1 −<br />

atom % 15 �<br />

× 100<br />

N excess (uninoculated plant)


102 R. Anand et al.<br />

In addition, strain P2b-2R was detected inside root, stem and needle<br />

tissue using a surface-sterilisation-trituration plating technique as well as<br />

with confocal laser scanning microscopy after insertion of green fluorescent<br />

protein into the bacterium. These results leave us with little doubt that pine<br />

can derive biologically significant amounts of nitrogen from endophytic<br />

diazotrophs.<br />

Two conclusions can be drawn from our work so far. Nitrogen-fixing bacteria<br />

can be isolated from internal root, stem and needle tissues of lodgepole<br />

pine seedlings and mature trees, some of which can contribute biologically<br />

significant amounts of fixed nitrogen to lodgepole pine seedlings under<br />

controlled environments. It is possible that a significant amount of plant<br />

nitrogen originates from diazotrophic endophytes in lodgepole pine, but<br />

additional research is required to elucidate the role of these bacteria in<br />

forest ecosystems.<br />

6.6<br />

Future Work<br />

We are currently involved in detecting in planta expression of strain P2b-2R<br />

nif genes to demonstrate that internal pine stem, root and needle tissues<br />

provide a suitable environment for nitrogen fixation, as we have already<br />

confirmed the existence of nif genes in this bacterial strain. These experiments<br />

will allow localisation of P2b-2R within the plant tissues as well<br />

as provide evidence that the observed nitrogen gains were in fact derived<br />

from the endophytic P2b-2R. In addition, we need to understand the effect<br />

of soil nitrogen availability on growth promotion by strain P2b-2R and the<br />

extent to which other mechanisms are responsible for growth promotion.<br />

Plant growth studies involving a non-diazotrophic mutant of strain P2b-<br />

2R, under various levels of available nitrogen will provide data to answer<br />

these questions.<br />

Systemically colonising endophytic bacteria such as P. polymyxa strain<br />

Pw2-RandP2b-2Rcouldalsobeusedasvectorstodeliverspecificgene<br />

products to plants. Such an approach may be more feasible than attempting<br />

to genetically alter the plant host directly. In addition, diazotrophic<br />

endophytic bacteria hold great potential for reducing application of fertilisers,<br />

especially of mineral N. Inoculation of forest seedlings with effective<br />

diazotrophic or plant growth promoting endophytic bacteria could<br />

enhance growth and yield of trees significantly, especially at nutrient-poor<br />

sites.However,thereismuchmoreworktobedoneifwearetounderstand<br />

these plant/microbe associations to the degree that we can manage<br />

them effectively for more efficient and sustainable nursery and field treatments.


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 103<br />

More knowledge of the population dynamics and activity of endophytic<br />

bacteriaintheirhostplantsisrequired.Aconsiderableresearcheffortisalso<br />

required to design strategies for the reinoculation of endophytic bacteria. In<br />

order to guarantee reproducibility, reliable methods of inoculum delivery<br />

should be developed. This is especially the case for the inoculation of<br />

trees with endophytic bacteria. Intense testing of different delivery systems<br />

has indicated that the efficacy of the application method for introducing<br />

endophytic bacteria into plant tissue is strain specific (Musson et al. 1995).<br />

The development of successful application technologies would fully depend<br />

on improving our understanding of how bacterial endophytes enter and<br />

colonise plants. This is just one aspect of the study of bacterial endophytes<br />

that needs to be undertaken in order to fully realise their potential use in<br />

forestry as in agriculture.<br />

Acknowledgements. The authors acknowledge the excellent help of Ms.S.E.<br />

Chanway in reference management and typing this manuscript.<br />

<strong>References</strong><br />

Achouak W, Normand P, Heulin T (1999) Comparative phylogeny of rrs and nif Hgenesin<br />

Bacillaceae. Int J Syst Bacteriol 49:961–967<br />

Ash C, Farrow JAE, Collins MD (1993) Molecular identification of rRNA group 3 bacilli<br />

using a PCR probe test: proposal for the creation of a genus Paenibacillus. Antonie van<br />

Leeuwenhoek 64:253–260<br />

Bal A (2003) Can lodgepole pine derive biologically significant amounts of N from bacterial<br />

endophytes? MSc Thesis, The University of British Columbia, Vancouver BC<br />

Balandreau J, Knowles R (1978) The rhizosphere. In: Dommergues YR, Krupa SV (eds) Interactions<br />

between non-pathogenic soil microorganisms and plants. Elsevier, Amsterdam,<br />

pp 243–268<br />

Baldani VLD, Döbereiner J (1980) Host plant specificity in the infection of cereals with<br />

Azospirillum ssp. Soil Biol Biochem 12:433–439<br />

Bent E, Chanway CP (1997) PGPR-mediated growth promotion effects on lodgepole pine<br />

can be inhibited by the presence of a rhizobacterial competitor. In: Ogoshi A et al. (eds)<br />

Plant growth-promoting Rhizobacteria: present status and future prospects. Nakanishi,<br />

Sapporo, pp 233–239<br />

Binkley D (1995) The influence of tree species on forest soils: processes and patterns.<br />

In: Mead DJ, Cornforth IS (eds) Proceedings of the trees and soil workshop, Lincoln<br />

University, 28 February–2 March 1994, Lincoln University Press, New Zealand, pp 1–33<br />

Boddey RM, Döbereiner J (1995) Nitrogen fixation associated with grasses and cereals:<br />

recent progress and perspectives for the future. Fertil Res 42:241–250<br />

Brock TD, Madigan MT, Martinko JM, Parker J (1994) Biology of microorganisms. Prentice<br />

Hall, New Jersey<br />

Brooks DS, Gonzalez CF, Appel DN, Filer TH (1994) Evaluation of endophytic bacteria as<br />

potential biological control agents for oak wilt. Biol Control 4:373–381<br />

Carroll G (1988) Fungal endophytes in stems and leaves: from latent pathogen to mutualistic<br />

symbiont. Ecology 69:2–9


104 R. Anand et al.<br />

Cavalcante VA, Döbereiner J (1988) A new acid-tolerant nitrogen-fixing bacterium associated<br />

with sugarcane. Plant Soil 108:23–31<br />

Chanway CP (1997) Inoculation of tree roots with plant growth promoting soil bacteria: an<br />

emerging technology for reforestation. For Sci 43:99–112<br />

Chanway CP, Holl FB (1991) Biomass increase and associative nitrogen fixation of mycorrhizal<br />

Pinus contorta seedlings inoculated with a plant growth promoting Paenibacillus<br />

strain. Can J Bot 69:507–511<br />

Chanway CP, Holl FB (1992) Influence of soil biota on Douglas-fir (Pseudotsuga menziesii<br />

(Mirb.) Franco) seedling growth: the role of rhizosphere bacteria. Can J Bot 70:1025–<br />

1031<br />

Chanway CP, Holl FB (1993) Field performance of spruce seedlings after inoculation with<br />

plant growth promoting rhizobacteria. Can J Microbiol 39:1084–1088<br />

Chanway CP, Holl FB (1994) Growth of outplanted lodgepole pine seedlings one year after<br />

inoculation with plant growth promoting rhizobacteria. For Sci 40:238–246<br />

Chanway CP, Shishido M, Holl FB (1994) Root-endophytic and rhizosphere plant growth<br />

promoting rhizobacteria for conifer seedlings. In: Ryder MH, Stephens PM, Bowen GD<br />

(eds) Improving plant productivity with rhizosphere bacteria. CSIRO Division of Soils<br />

1994:72–74<br />

Chanway CP, Shishido M, Jungwirth S, Nairn J, Markham G, Xiao, Holl FB (1997) Second<br />

year growth responses of outplanted conifer seedlings inoculated with PGPR. In:<br />

Ogoshi A, Kobayashi K, Homma Y, Kodama F, Kondo N, Akino S (eds) Plant growthpromoting<br />

Rhizobacteria present status and future prospects. Nakanishi, Sapporo,<br />

pp 172–176<br />

Chanway CP, Shishido M, Nairn J, Jungwirth S, Markham J, Xiao G, Holl FB (2000) Endophytic<br />

colonisation and field responses of hybrid spruce seedlings after inoculation<br />

with plant growth promoting rhizobacteria. For Ecol Manage 133:81–88<br />

Clay K (1988) Fungal endophytes of grasses: a defensive mutualism between plants and<br />

fungi. Ecology 69:10–16<br />

Döbereiner J, Urquiaga S, Boddey RM (1995) Alternatives for nitrogen nutrition of crops in<br />

tropical agriculture. Fertil Res 42:339–346<br />

Fitter AH, Garbaye J (1994) Interaction between mycorrhizal fungi and other soil organisms.<br />

Plant Soil 159:123–132<br />

Glick BR (1995) The enhancement of plant growth by free-living bacteria. Can J Microbiol<br />

41:109–117<br />

Gardner JM, Feldman AW, Zablotowicz M (1982) Identity and behavior of xylem-residing<br />

bacteria in rough lemon roots of Florida citrus trees. Appl Environ Microbiol 43:1335–<br />

1342<br />

Hallmann J, Quadt-Hallmann A, Mahafee WF, Kloepper JW (1997) Bacterial endophytes in<br />

agricultural crops. Can J Microbiol 43:895–914<br />

Holl FB, Chanway CP, Turkington R, Radley RA (1988) Response of crested wheatgrass<br />

(Agropyron cristatum L), perennial ryegrass (Lolium perenne L) and white clover (Trifolium<br />

repens L.) to inoculation with Paenibacillus polymyxa. Soil Biol Biochem 20:19–24<br />

Hollis JP (1951) Bacteria in healthy potato tissue. Phytopathology 41:320–366<br />

Hunt R (1982) Plant growth curves. University Park Press, Baltimore<br />

Hurek T, Reinhold-Hurek B, Van Montagu M, Kellenberger E (1994) Root colonisation and<br />

systemic spreading of Azoarcus sp. strain BH72 in grasses. J Bacteriol 176:1913–1923<br />

James EK (2000) Nitrogen fixation in endophytic and associative symbiosis. Field Crops Res<br />

65:197–209<br />

James EK, Olivares FL (1997) Infection and colonisation of sugar cane and other graminaceous<br />

plants by endophytic diazotrophs. Crit Rev Plant Sci 17:77–119


6 Research on Endophytic Bacteria: Recent Advances with Forest Trees 105<br />

Kloepper JW (1993) Plant growth-promoting rhizobacteria as biological control agents. In:<br />

Metting FB (ed) Soil microbial ecology applications in agricultural and environmental<br />

management. Dekker, New York, pp 255–274<br />

Kloepper JW, Lifshitz R, Zablotowicz RM (1989) Free living bacterial inocula for enhancing<br />

crop productivity. Trends Biotechnol 7:39–44<br />

Law R, Lewis DH (1983) Biotic environments and the maintenance of sex – some evidence<br />

from mutualistic symbioses. Biol J Linn Soc 20:249–276<br />

Lehninger AL (1975) Biochemistry: the molecular basis of cell structure and function.<br />

Worth, New York<br />

Li CY, Massicotte HB, Moore LV (1992) Nitrogen fixing Bacillus sp. associated with Douglas<br />

fir tuberculate ectomycorrhizae. Plant Soil 140:35–40<br />

Lodewyckx C, Vangronsveld I, Porteous F, Moore ERB, Taghavi S, Mezgeay M, Vander<br />

lella D (2002) Endophytic bacteria and their potential applications. Crit Rev Plant Sci<br />

21:583–606<br />

Mavingui P, Laguerre PG, Berge O, Heulin T (1992) Genotypic and phenotypic variability of<br />

Paenibacillus polymyxa in soil and in the rhizosphere of wheat. Appl Environ Microbiol<br />

58:1894–1903<br />

McInroy JA, Kloepper JW (1994) Novel bacterial taxa inhabiting internal tissue of sweet corn<br />

and cotton. In: Ryder MH, Stephens PM, Bowen GD (eds) Improving plant productivity<br />

with rhizosphere bacteria. CSIRO, Melbourne, Australia<br />

Mengoni A, Mocali S, Surico G, Tegli S, Fani R (2003) Fluctuation of endophytic bacteria<br />

and phytoplasmosis in elm trees. Microbiol Res 158:363–369<br />

Misaghi IJ, Donndelinger CR (1990) Endophytic bacteria in symptom-free cotton plants.<br />

Phytopathology 80:808–811<br />

Mocali S, Bertelli E, Di Cello F, Mengoni A, Sfalanga A, Viliani F, Caciotti A, Tegli S, Surico G,<br />

Fani R (2003) Fluctuation of bacteria isolated from elm tissues during different seasons<br />

and from different plant organs. Microbiol Res 154:105–114<br />

Musson G, McInroy JA, Kloepper JW (1995) Development of delivery systems for introducing<br />

endophytic bacteria into cotton. Biocontrol Sci Technol 5:407–416<br />

O’Neill GA, Chanway CP, Axelrood PE, Radley RA, FB Holl (1992) Growth response<br />

specificity of spruce inoculated with coexistent rhizosphere bacteria. Can J Bot<br />

70:2347–2353<br />

Paul EA, Clark FE (1989) Soil microbiology and biochemistry. Academic, New York<br />

Pokojska-Burdziej A (1982) The effect of microorganisms, microbial metabolites and plant<br />

growth regulators on the growth of pine seedlings (Pinus sylvestris L.). Pol J Soil Sci<br />

15:137–143<br />

Rennie RJ (1981) A single medium for the isolation of acetylene-reducing (dinitrogen-fixing)<br />

bacteria from soils. Can J Microbiol 27:8–14<br />

Rennie RJ, Rennie DA, Fried M (1978) Concepts of 15 N usage in dinitrogen fixation studies.<br />

In: Isotopes in biological dinitrogen fixation. International Atomic Energy Agency,<br />

Vienna, pp 107–133<br />

Reinhold-Hurek B, Hurek T (1998) Interactions of Gramineous plants with Azoarcus spp. and<br />

other diazotrophs: identification, localization, and perspectives to study their function.<br />

Crit Rev Plant Sci 17:29–54<br />

Rhodes-Roberts M (1981) The taxonomy of some nitrogen fixing Paenibacillus species<br />

with special reference to nitrogen fixation. In: Berkeley RCW, Goodfellow M (eds)<br />

The aerobic-endosperm forming bacteria classification and identification. Academic,<br />

London, pp 315–335<br />

Schardl CL, Leuchtmann A, Spiering MJ (2004) Symbioses of grasses with seedborne fungal<br />

endophytes. Annu Rev Plant Biol 55:315–340


106 R. Anand et al.<br />

Shishido M (1997) PGPR for interior spruce seedlings. PhD Thesis University of British<br />

Columbia, Vancouver BC<br />

Shishido M, Chanway CP (2000) Colonisation and growth promotion of outplanted spruce<br />

seedlings pre-inoculated with plant growth-promoting rhizobacteria in the greenhouse.<br />

Can J For Res 30:845–854<br />

Shishido M, Loeb BM, Chanway CP (1995) External and internal root colonisation of lodgepole<br />

pine seedlings by two growth-promoting Paenibacillus strains originated from<br />

different root microsites. Can J Microbiol 41:707–713<br />

Shishido M, Massicotte HB, Chanway CP (1996) Effect of plant growth promoting Paenibacillus<br />

strains on pine and spruce seedling growth and mycorrhizal infection. Ann Bot<br />

77:433–441<br />

Sprent JL, James EK (1995) N2-fixation by endophytic bacteria: questions of entry and operation.<br />

In: Fendrick I (ed) NATO ASI series, Azospirillum VI and related microorganisms,<br />

vol G. Springer, Berlin Heidelberg New York, pp 15–30<br />

Tervet IW, Hollis JP (1948) Bacteria in the storage organs of healthy plants. Phytopathology<br />

38:960–967<br />

TimmuskS,Wagner EGH(1999) The plant-growth-promotingrhizobacterium Paenibacillus<br />

polymyxa induces changes in Arabidopsis thaliana gene expression: a possible connection<br />

between biotic and abiotic stress responses. Mol Plant-Microbe Interact 12:951–959<br />

Triplett EW (1996) Diazotrophic endophytes: progress and prospects for nitrogen fixation<br />

in monocots. Plant Soil 186:29–38<br />

Whitesides SK, Spotts RA (1991) Frequency, distribution, and characteristics of endophytic<br />

Pseudomonas syringae in pear trees. Phytopathology 81:453–457<br />

Wilson D (1995) Endophyte: the evolution of a term, and clarification of its use and definition.<br />

Oikos 73:274–276<br />

Zak B (1971) Characterization and classification of mycorrhizae of Douglas fir. II. Pseudotsuga<br />

menziesii + Rhizopogon vinicolor. Can J Bot 49:1079–1084


7<br />

Biodiversity of Fungal Root-Endophyte<br />

Communities and Populations,<br />

in Particular of the Dark Septate Endophyte<br />

Phialocephala fortinii s. l.<br />

Thomas N. Sieber, Christoph R. Grünig<br />

7.1<br />

Introduction<br />

The peripheral root tissues form a morphologically, physically and chemically<br />

complex microcosm that provides different habitats for diverse communities<br />

of microorganisms. This microcosm is not stable, and changes<br />

over space and time because the boundaries between soil, rhizosphere, and<br />

living roots are continually shifted as a result of root growth and the constant<br />

modification of nearby soil by root mechanical and metabolic activity<br />

(Foster et al. 1983). Microorganisms colonise the rhizoplane, epidermis and<br />

outer cortex in a nonrandom patchy manner and contribute to the modification<br />

of the soil-rhizosphere-root continuum. Microorganisms affect<br />

their plant hosts, and hosts reciprocally affect their symbionts, leading to<br />

a feedback that drives changes in both the microbial and plant communities<br />

(Bever et al. 1997). Many soil bacteria and fungi are able to colonise epidermal<br />

and outer cortical cells of healthy roots inter- and intra-cellularly.<br />

A comparatively small number of organisms, e.g. mycorrhizal fungi, endophytic<br />

and pathogenic fungi and bacteria, possess, however, the ability<br />

to cross the inner boundary of the rhizosphere and to penetrate deeper<br />

into the root (Bazin et al. 1990). The interaction of host and endophyte<br />

depends on the disposition of host and fungus or bacterium and the environmental<br />

conditions, but may be neutral, mutualistic or antagonistic and<br />

may change over time. Some endophytic fungi adopt mycorrhizal functions<br />

and/or place plants at a competitive advantage against herbivores, insect<br />

pests or pathogens (Carroll 1988; Hawksworth 1991). Other endophytes<br />

can switch to a pathogenic behaviour when conditions are unfavourable<br />

for the host (Schulz et al. 1999). The biodiversity of root endophyte communities<br />

varies in relation to environmental factors, type of vegetation,<br />

Thomas N. Sieber: Swiss Federal Institute of Technology, Department of Environmental<br />

Sciences, Institute of Integrative Biology, Forest Pathology and Dendrology, 8092 Zürich,<br />

Switzerland, E-mail: thomas.sieber@env.ethz.ch<br />

Christoph R. Grünig: Swiss Federal Institute of Technology, Department of Environmental<br />

Sciences, Institute of Integrative Biology, Forest Pathology and Dendrology, 8092 Zürich,<br />

Switzerland<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


108 T.N. Sieber, C.R. Grünig<br />

spatiotemporal patterns of the root microcosm and interactions among<br />

microorganisms. There is currently an urgent need to assess biodiversity<br />

in pristine ecosystems and to use these data as references to measure the<br />

effects of disturbances on diversity and to better enable informed decisionmaking<br />

on the fate of threatened natural habitats (Cannon 1997). Threats<br />

may come from a variety of sources, including exploitation by logging,<br />

machine-graded soils, urban development, pollution, climate change and<br />

input of pesticides and fertilisers. Biodiversity can be explored at several<br />

levels, i.e. in terms of communities, species and populations (Hawksworth<br />

1991). Here, we will explore current knowledge on the biodiversity of nonmycorrhizal<br />

fungal root endophytes at all levels. The first part of this review<br />

will be dedicated to biodiversity at the community level in relation to environmental<br />

factors. In the second part, special emphasis will be placed on the<br />

diversity of dark septate endophytes (DSE), in particular of Phialocephala<br />

fortinii s. l.<br />

Readers of this chapter should always bear in mind that the methods of<br />

detection are highly selective and, thus, the species list and species diversity<br />

derived for any habitat will be incomplete and will be biased in respect to<br />

physiological features selected for by the method used [Sieber 2002; Swift<br />

1976; see Chaps. 9 (Bayman and Otero), 18 (Bloemberg and Camacho<br />

Carvajal) and 19 (Van Overbeek et al.)].<br />

7.2<br />

Species Diversity of Root Endophyte Communities<br />

“Species diversity” comprises two distinct components: the total number of<br />

species, which ecologists refer to as “species richness”, and “evenness” or<br />

equitability, which refers to how species abundances are distributed among<br />

the species present. An ecosystem is said to be more diverse if many species<br />

with equal population sizes are present and less diverse if some species<br />

are rare and a few are very common. Other helpful terms are “spectrum<br />

of species” or “community composition” to describe habitat or ecosystem<br />

differences with respect to the species found. The species diversity and the<br />

species spectrum of root-endophyte communities are related to various<br />

factors, which can tentatively be arranged into four groups: (1) geography<br />

and climate, (2) soil, (3) multitrophic interactions, and (4) natural and<br />

anthropogenic disturbances. This grouping is rather artificial and does<br />

not account for the intricate interplay among factors that often makes it<br />

impossible to determine the contribution of each factor. Another aspect<br />

obscuringtheeffectsofdifferentfactorsisthatofsitehistory,i.e.the<br />

dynamics of plant and endophyte communities. Nevertheless, the above<br />

grouping seems to be the most appropriate structure for this section.


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 109<br />

7.2.1<br />

Geography and Climate<br />

Fungal species diversity is higher in tropical than in temperate regions<br />

owinginparttothegreatdiversityofhosts,butalsototheoptimalgrowth<br />

conditions for many fungi as a result of the hot and moist climate (Cannon<br />

and Hawksworth 1995). Whether this relationship is also valid for fungal<br />

root endophytes remains to be tested. Compared to habitats in the temperate<br />

or the tropical zones, species diversity is distinctly reduced in arctic-alpine<br />

environments, not only because of the lower number of available<br />

host species, but also with respect to the number of endophyte species in<br />

each host. For example, only seven root endophyte species were detected in<br />

Dryas octopetala in arctic Spitsbergen (Fisher et al. 1995) [Table 7.1(i)]. Correspondingly,<br />

species richness in Erica carnea was highest at an altitude of<br />

640 m and lowest at 2,140 m in Switzerland (Oberholzer-Tschütscher 1982).<br />

The species spectra differed greatly among sites, as expressed by very low<br />

between-site similarities [Table 7.1(ii)]. Evenness was lowest at the lowest<br />

altitude where the comparatively species-rich community was dominated<br />

by only four to five species.<br />

There is strong evidence for a shift from arbuscular mycorrhizal fungi<br />

(AMF) and ectomycorrhizal fungi (ECM) in temperate habitats towards<br />

symbioses of uncertain status, especially dark septate endophytes (DSE),<br />

in arctic-alpine ecosystems (Bledsoe et al. 1990; Christie and Nicolson<br />

1983; Read and Haselwandter 1981; Väre et al. 1992). Correspondingly, the<br />

frequency of roots colonised by Phialocephala fortinii s. l., a ubiquitous and<br />

dominant DSE in conifer roots (see Sect. 7.3.3), was positively correlated<br />

with the altitude in forest ecosystems (Ahlich and Sieber 1996).<br />

Weather and climatic conditions are assumed to have a weaker direct<br />

effect on species diversity of endophyte assemblages in root tissues than in<br />

aerial plant parts due to the insulating and compensating properties of soils<br />

(Fitter et al. 1985). Thus, changes in root endophyte assemblages become<br />

manifest only if the climatic conditions deviate from the “normal” over<br />

an extended period of time, i.e. if the climate changes. In fact, long-term<br />

changes in mean annual temperature, frequency and amount of precipitation,aswellasenhancedCO2<br />

may affect root endophyte diversity through<br />

shifts in the quantity and quality of photosynthates and secondary plant<br />

metabolites translocated to the roots, the rate of root turnover, and shifts in<br />

the competivity of endophytes and other soil microorganisms (Coûteaux et<br />

al. 1999; Rillig et al. 1999; Körner 2000). However, nothing is known about<br />

the direction and magnitude of effects on root-endophyte diversity.


110 T.N. Sieber, C.R. Grünig<br />

Table 7.1. Influence of geographical, physical, chemical and biological factors on species diversity and similarity of communities of fungal root<br />

endophytes<br />

Reference<br />

Pairwise<br />

similarities of<br />

Sample<br />

sizee communities f<br />

Total<br />

number<br />

of<br />

isolates<br />

Evenness<br />

indexd Number<br />

of very<br />

abundant<br />

speciesc Adjusted<br />

species<br />

richnessb Hosts and factors Observed<br />

species<br />

richnessa (2) (3) (4) (5)<br />

(i) Dryas octopetala R Fisheretal.<br />

(1) Spitzbergen, site A 4 3.8 ± 0.4 2.4 0.81 42 50 0.73<br />

1995<br />

(2) Spitzbergen, site B 7 5.8 ± 0.4 3.8 0.83 24 50 –<br />

(ii) Erica carnea R Oberholzer-<br />

(1) Fläsch (640 m) 35 26.5 ± 2.2 4.2 0.51 296 333 0.26 0.21 0.30 Tschütscher<br />

(2) Näfels (760 m) 19 17.1 ± 1.2 4.7 0.71 219 120 – 0.20 0.19 1982<br />

(3) Davos-Wolfgang (1,640 m) 22 21.8 ± 0.4 7.0 0.72 173 298 – – 0.24<br />

(4) Davos-Schatzalp (2,140 m) 12 10.6 ± 1.0 2.7 0.68 235 300 – – –<br />

(iii) Picea abies R Kattnerand<br />

(1) Soil pH neutral 27 26.9 ± 0.3 11.9 0.70 153 480 0.57<br />

Schönhar<br />

(2) Soil pH acidic 29 28.8 ± 0.4 12.6 0.72 154 480 –<br />

1990<br />

(iv) Alnus glutinosa R Fisheretal.<br />

(1) Submerged roots 45 44.1 ± 0.8 17.4 0.66 114 40 0.37<br />

1991<br />

(2) Non-submerged roots 31 29.2 ± 1.2 13.8 0.75 126 40 –<br />

(v) Rhizophora mucronata R Ananda and<br />

(1) Low-tide level 6 5.3 ± 0.7 2.2 0.80 21 30 0.50 0.38<br />

Sridhar 2002<br />

(2) Mid-tide level 14 9.7 ± 1.2 3.6 0.87 34 30 – 0.58<br />

(3) High-tide level 10 9.4 ± 0.6 3.2 0.91 17 30 – –


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 111<br />

Table 7.1. (continued)<br />

Reference<br />

Pairwise<br />

similarities of<br />

Sample<br />

sizee communities f<br />

Total<br />

number<br />

of<br />

isolates<br />

Evenness<br />

indexd Number<br />

of very<br />

abundant<br />

speciesc Adjusted<br />

species<br />

richnessb Hosts and factors Observed<br />

species<br />

richnessa (2) (3) (4) (5)<br />

(vi) Phragmites australis Wirsel et al.<br />

Location 1 R<br />

2001<br />

(1) Flooded site 10 9.5 ± 0.7 4.6 0.74 63 45 0.52 0.52 –<br />

(2) Dry site 13 12.9 ± 0.3 7.4 0.80 51 45 – – 0.75<br />

Location 2<br />

(3) Flooded site 13 11.9 ± 0.9 5.6 0.71 47 45 0.58<br />

(4) Dry site 11 10.0 ± 0.9 4.8 0.72 52 45 –<br />

(vii) Triticum aestivum Sieber et al.<br />

Development stage: R<br />

1988<br />

(1) One leaf – end of tillering 63 61.0 ± 1.3 15.3 0.59 547 5040 0.67<br />

(2) Inflorescence emerged – caryopsis hard 62 39.5 ± 2.9 6.3 0.53 1857 3360 –<br />

(viii) Triticum aestivum Sieber et al.<br />

Preceding crop: R<br />

1988<br />

(1) Sugar beet 51 44.7 ± 2.1 9.3 0.56 623 2400 0.58 0.57 0.67<br />

(2) Red clover 49 48.3 ± 3.3 10.9 0.60 413 1200 – 0.60 0.59<br />

(3) Maize 44 35.9 ± 2.2 4.7 0.45 773 2400 – – 0.65<br />

(4) Potatoes 39 33.5 ± 1.9 6.9 0.61 595 2400 – – –


112 T.N. Sieber, C.R. Grünig<br />

Table 7.1. (continued)<br />

Reference<br />

Pairwise<br />

similarities of<br />

Sample<br />

sizee communities f<br />

Total<br />

number<br />

of<br />

isolates<br />

Evenness<br />

indexd Number<br />

of very<br />

abundant<br />

speciesc Adjusted<br />

species<br />

richnessb Hosts and factors Observed<br />

species<br />

richnessa (2) (3) (4) (5)<br />

(ix) Various vegetables R Narisawa<br />

(1) Eggplant 7 5.5 ± 0.9 3.2 0.71 35 45 0.67 0.55 0.67 0.86 et al. 2002<br />

(2) Tomato 5 4.7 ± 0.4 2.7 0.78 19 45 – 0.44 0.60 0.50<br />

(3) Melon 4 3.9 ± 0.3 2.6 0.83 17 45 – – 0.44 0.55<br />

(4) Strawberry 5 4.5 ± 0.6 2.1 0.71 21 45 – – – 0.67<br />

(5) Chinese cabbage 7 5.9 ± 0.8 4.4 0.79 29 45 – – –<br />

(x) Betula pendula T Görke<br />

(1) Plantation in cleared windthrow 30 12.5 ± 1.7 10.6 0.59 87 160 0.34 0.29 0.38 (1998)<br />

(2) Natural regeneration<br />

11 9.3 ± 1.0 5.7 0.73 27 75 – 0.41 0.50<br />

in untouched windthrow<br />

(3) Natural regeneration<br />

18 10.9 ± 1.5 7.4 0.60 48 100 – – 0.57<br />

in cleared windthrow<br />

(4) Natural regeneration<br />

in low density forestg 17 9.3 ± 1.5 6.3 0.70 58 100 – – –<br />

(x) Pinus sylvestris T Görke<br />

(1) Plantation in cleared windthrow 16 6.2 ± 1.2 6.8 0.72 56 160 0.42 0.40 0.38 (1998)<br />

(2) Natural regeneration<br />

8 5.5 ± 1.0 4.7 0.79 20 75 – 0.55 0.53<br />

in untouched windthrow<br />

(3) Natural regeneration<br />

14 7.0 ± 1.2 6.9 0.69 26 100 – – 0.38<br />

in cleared windthrow<br />

(4) Natural regeneration in low density forest 7 5.4 ± 0.9 4.8 0.80 19 100 – – –


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 113<br />

Table 7.1. (continued)<br />

Reference<br />

Pairwise<br />

similarities of<br />

Sample<br />

sizee communities f<br />

Total<br />

number<br />

of<br />

isolates<br />

Evenness<br />

indexd Number<br />

of very<br />

abundant<br />

speciesc Adjusted<br />

species<br />

richnessb Hosts and factors Observed<br />

species<br />

richnessa (2) (3) (4) (5)<br />

(xi) Erica carnea R Cevniket<br />

(1) Control (Cd 1.4; Pb 171; Zn 61.8) al. 2000<br />

h 9 8.2 ± 0.7 3.9 0.78 104 240 0.50 0.60 0.63<br />

(2) Low pollution (Cd 6.9; Pb 667; Zn 177) 7 6.9 ± 0.2 3.3 0.81 72 240 – 0.67 0.71<br />

(3) High pollution (Cd 35.8; Pb 5422; Zn 582) 11 10.5 ± 0.6 8.0 0.87 107 240 – – 0.67<br />

(4) Highest pollution<br />

10 8.9 ± 0.8 3.8 0.72 142 240 – – –<br />

(Cd 87.7; Pb 31320; Zn 1330)<br />

a Diversity index N0 according to Hill (1973); number of species<br />

b Mean and standard error of the number of species adjusted<br />

to the lowest within-study number of isolates using rarefaction according to Hurlbert (1971)<br />

c Diversity index N2 according to Hill (1973)<br />

d Evenness index according to Hill (1973); this index converges towards 1 as one species tends to dominate<br />

e Number of root segments (R) or trees (T) examined<br />

f Soerensen index (Soerensen 1948); column numbers (in brackets) correspond to factor identifiers in the column “Hosts and factors”;<br />

0 ≤ Soerensen index ≤ 1, the index is 0 if two communities have no species in common, and it is 1 if all species occur in both communities<br />

g Low density forest = selectively logged forest stand; aim: increased solar radiation within the stand<br />

h Concentrations of heavy metals in micrograms per gram of soil


114 T.N. Sieber, C.R. Grünig<br />

7.2.2<br />

Soil<br />

Soil and rhizosphere are highly variable habitats. Chemical properties such<br />

as pH or the availability of minerals and carbohydrates may vary significantly<br />

within a few centimetres of soil (Papritz and Flühler 1991). Similarly,<br />

differences in soil texture and water regime contribute to the variability<br />

of soils. In addition, roots constantly modify the nearby soil structure by<br />

depletion of minerals, ions and water and by the secretion of root exudates.<br />

Soils offer habitats for various communities of microorganisms including<br />

potential root endophytes. Plant and microbial metabolites may differentially<br />

influence the surrounding soil and change some of its properties, thus<br />

preparing the soil for the microorganisms of the next successional stage<br />

(Van Der Putten 2003).<br />

Physical and Chemical Soil Characteristics<br />

SoilpHhadaneffectoncommunitycompositionbutnotonspeciesdiversity<br />

of endophytic fungi in Norway spruce roots (Picea abies) (Kattner<br />

and Schönhar 1990) [Table 7.1(iii)]. The similarity of only 57% of the endophyte<br />

communities in roots from neutral and acidic soils reflects either<br />

the selectivity of soil pH or the historical presence/absence of certain endophyte<br />

species, e.g. endophytes with low dispersion and/or survival rates.<br />

For example, Phialocephala fortinii preferentially occurs in roots growing<br />

in acidic soils (Ahlich et al. 1998).<br />

Species richness was not related to soil texture in wheat roots (Triticum<br />

aestivum) (Riesen and Sieber 1985; Sieber et al. 1988). However, texture affected<br />

the frequency of Microdochium bolleyi and Periconia macrospinosa.<br />

M. bolleyi was more frequently isolated from roots originating from silty<br />

loam, whereas P. macrospinosa was isolated more often from roots growing<br />

in pure loam.<br />

Root endophytes differ in their ability to metabolise minerals and carbohydrates,<br />

making some endophytes more successful than others in a given<br />

habitat. DSE are thought to be excellent metabolisers of phosphorus (P)<br />

and to mediate P uptake for their hosts (Jumpponen et al. 1998; Barrow<br />

and Osuna 2002). In fact, DSE were more abundant in habitats poor in<br />

P (Haselwandter and Read 1982; Ruotsalainen et al. 2002). Similarly, differential<br />

utilisation of carbohydrates as well as which carbohydrates were<br />

available determined fungal species diversity and endophyte-community<br />

composition in the experiments of Hadacek and Kraus (2002).


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 115<br />

Water Regime<br />

The water regime in soils and streams has a strong impact on species diversity<br />

and especially on the species spectrum of endophytic fungi (see<br />

Chap. 10 by Bärlocher). In roots of the same tree, 45 species were isolated<br />

from roots submerged in a river as opposed to only 31 species from nonsubmerged<br />

roots (Fisher et al. 1991) [Table 7.1(iv)]. The similarity of the<br />

community composition in submerged and non-submerged roots of the<br />

same individual black-alder tree was only 37%. Colonisation of submerged<br />

roots by aquatic hyphomycetes, together with the absence or scarcity of<br />

these specialists in non-submerged roots, emphasise the importance of the<br />

milieu in which roots grow in determining the composition and diversity<br />

of endophyte communities. For example, high water tables restricted the<br />

occurrence of P. fortinii in wetlands (Addy et al. 2000). The endophyte<br />

species diversity in roots of the mangrove Rhizophora mucronata strongly<br />

depended on the tidal level at which the roots were collected. Diversity<br />

was highest at the mid-tide level, i.e. the zone submerged in seawater approximately<br />

half of the time, and roots from the high-tide and the low-tide<br />

level had, on average, only 38% of species in common (Ananda and Sridhar<br />

2002) [Table 7.1(v)]. Flooding and site conditions affected endophyte<br />

species spectra but not species richness in roots of common reed (Phragmites<br />

australis) (Wirsel et al. 2001) [Table 7.1(vi)]. In contrast, species<br />

spectra in bracken rhizomes (Pteridium aquilinum) did not differ among<br />

wetland and woodland sites (Petrini et al. 1992).<br />

7.2.3<br />

Multitrophic Interactions<br />

The diversity of soil microorganisms is tremendous; 1 g soil can contain<br />

between 5,000 and 10,000 species of microorganisms (Torsvik et al. 1990).<br />

However, only 1,200 species of fungi have been isolated from soil (Watanabe<br />

1994), perhaps because, as estimates suggest, only 17% of known fungi can<br />

be readily grown in culture (Hawksworth 1991). If this percentage were<br />

applied to the 1,200 species as suggested by Watanabe (1994), this would<br />

give an estimate of approximately 7,000 species of soil fungi (Bridge and<br />

Spooner 2001). The total length of fungal hyphae varies greatly according<br />

to soil type and soil biology and has been reported to be as high as 66,900 m<br />

in 1 g dry soil (Bååth and Söderström 1979). The high number of species<br />

and the high amount of microbial biomass in such small volumes of soil<br />

suggest that multitrophic interactions among soil bacteria, soil fungi, soil<br />

microfauna and plants are frequent. Interspecific competition may be “the”<br />

factor that overrides all others in regulating species abundance of soil fungi<br />

(Gochenaur 1984). If a community is dominated by inter- and intra-specific


116 T.N. Sieber, C.R. Grünig<br />

competition, the resources are more likely to be fully exploited. Endophyte<br />

speciesdiversityandspectrumwillthendependontherangeofavailable<br />

resources, including host tissues, the extent to which species are specialists,<br />

antagonism among competitors, their ability to overcome host defences and<br />

the permitted extent of habitat overlap.<br />

Microdochium bolleyi is a frequent and successful endophyte in cereal<br />

roots, where it functions as an effective antagonist of various root<br />

pathogens. For example, its presence in wheat roots was negatively correlated<br />

with the presence of Septoria nodorum, the causal agent of glume<br />

blotch disease of wheat (Riesen and Sieber 1985; Sieber et al. 1988). Similarly,<br />

M. bolleyi inhibited various Fusarium species and Gaeumannomyces<br />

graminisvar. tritici (KirkandDeacon1987;Reinecke1978).Whether M. bolleyi<br />

interacts with these pathogens indirectly by inducing systemic resistance<br />

in the host plant, or directly by either parasitising pathogens or<br />

producing inhibitory metabolites, remains to be examined.<br />

Thephenologicalstateoftherootsand/ortheseasonmayinfluenceendophyte<br />

species diversity by affecting the probability of interactions among<br />

endophytic thalli. For example, the number of dominant species was higher<br />

in young than in mature winter wheat, presumably because freshly established<br />

thalli were small. Growth was reduced due to the cold temperatures<br />

in winter, making hyphal interference less likely and/or weaker and, thus,<br />

also allowing less competitive fungi or fungi better adapted to cold temperatures<br />

to establish endophytic thalli [Table 7.1(vii)] (Riesen and Sieber<br />

1985; Sieber et al. 1988). This situation changed in spring and summer,<br />

when the growth rate of endophytic thalli increased, making intra- and<br />

inter-species hyphal interactions more probable, leading to the dominance<br />

of the few most competitive species.<br />

Similar to mycorrhiza, strict host specificity is the exception rather than<br />

the rule for fungal root endophytes (Bruns et al. 2002; Jumpponen et al.<br />

2004). However, the likelihood of occurrence of some endophyte species<br />

increases in the presence of particular host species, suggesting fungal host<br />

preference or shared habitat preferences. The diversity of the plant community<br />

in which the host species grows may, therefore, influence rootendophyte<br />

diversity similarly as it has been shown to affect diversity of soil<br />

microfungi (Christensen 1981, 1989). Ahlich and Sieber (1996) presented<br />

an example of the importance of the plant community in determining the<br />

spectra of fungi associated with the host. The dominant root endophytes<br />

of European beech (Fagus sylvatica), Cryptosporiopsis radicicola and Cylindrocarpon<br />

didymum, were rare or absent in roots of Scots pine (Pinus<br />

sylvestris) growing in monoculture. Likewise, P. fortinii,thedominantroot<br />

endophyte of Scots pine, was rare or absent in monocultures of beech.<br />

However, when the roots originated from mixed stands of Scots pine and<br />

beech, Scots pine roots showed a comparatively high rate of colonisation


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 117<br />

by C. radicicola and C. didymum. Correspondingly, the roots of beech were<br />

frequently colonised by P. fortinii in mixed stands. In contrast, frequency<br />

of colonisation of Betula papyrifera and Pseudotsuga menziesii seedlings<br />

by DSE was not affected by whether or not the plants were grown in mixed<br />

culture or in monoculture (Jones et al. 1997).<br />

In agriculture, the preceding crop may significantly affect endophyte diversity<br />

of the current crop. For example, species richness and the number<br />

of dominant species were significantly higher when wheat (Triticum aestivum)<br />

followed red clover than when it followed potatoes [Table 7.1(viii)]<br />

(Sieber et al. 1988). On average, only 59% of the endophyte species were indifferent<br />

to whether the preceding crop was clover or tomatoes. The range<br />

of indifferent endophyte species lay between 57% and 67% for other pairs<br />

of preceding crops [Table 7.1(viii)]. This observation may be related to<br />

differences in the spectra of endophytes that had colonised the preceding<br />

crop. Specific secondary metabolites and debris produced by the preceding<br />

crop, as well as the type and amount of agrochemicals (fertilisers, biocides,<br />

leafage killers) applied to the preceding crops may be other factors<br />

influencing both diversity and stimulation/inhibition of endophytes.<br />

When different vegetables are grown in the same soil, some endophytehost<br />

associations occur more frequently than others, suggesting host preference<br />

or adaptation. The similarity of the spectra of endophyte species<br />

among host species was as low as 44% in an experiment performed by Narisawa<br />

et al. (2002) [Table 7.1(ix)]. It is not known whether plants are able<br />

to actively recruit endophytes and vice-versa. Plant defence compounds<br />

probably select for certain rhizosphere microorganisms. Some evidence<br />

for such mechanisms comes from nematode and mycorrhiza research. Secondary<br />

metabolites released by roots of Thuja occidentalis upon attack by<br />

weevil larvae attracted entomopathogenic nematodes (Van Tol et al. 2001).<br />

Dormant propagules of mycorrhizal fungi were stimulated to germinate<br />

by chemical messengers from the host (Bruns et al. 2002). Correspondingly,<br />

mycelia of AMF were inhibited by non-host metabolites (Oba et al.<br />

2002). Nothing is known about whether certain root endophytes release<br />

“pheromones” to attract roots of host plants.<br />

7.2.4<br />

Natural and Anthropogenic Disturbances<br />

Anthropogenic and natural disturbances affect the species spectrum of<br />

plant communities and consequently also the communities of cohabiting<br />

microorganisms. Forest-management practices such as planting of trees,<br />

selective cutting or clearing of windthrows had a distinct effect on the<br />

endophytic mycobiota in the roots of forest trees (Görke 1998). Maximally


118 T.N. Sieber, C.R. Grünig<br />

42% of the endophyte species were common to both planted and naturally<br />

regenerated trees [Table 7.1(x)]. Considering naturally regenerated trees<br />

only, species richness and the number of dominant species was highest<br />

in the cleared windthrow. Probably, endophyte diversity and community<br />

composition would also change as a consequence of gap formation by man<br />

and/or wind storm, which eliminates some hosts but creates habitats for<br />

many other hosts, i.e. ruderal plant species.<br />

Mycorrhization and root-endophyte colonisation of naturally regenerated<br />

seedlings of Betula platyphylla var. japonica in soils of machine-graded<br />

ski slopes depended on the time elapsed since disturbance (Hashimoto<br />

and Hyakumachi 2000). Seedlings thrived well only in soil samples from<br />

soils disturbed more than 3 years previously and mycorrhization was significantly<br />

higher in these samples. In contrast, colonisation of roots by<br />

DSE was distinctly higher in seedlings sampled from soils disturbed only<br />

1–3 years before sampling. In another study, the majority of naturally established<br />

seedlings of bishop pine (Pinus muricata)werecolonisedbyDSE<br />

shortly after wildfire, indicating that a resident inoculum (chlamydospores,<br />

microsclerotia) survived the fire (Horton et al. 1998). Species richness of<br />

endophytes in roots of Erica carnea was highest at sites where soil pollution<br />

by heavy metals was high, but DSE occurred less frequently in the<br />

heavily polluted soils (Cevnik et al. 2000) [Table 7.1(xi)]. Endophytic fungi<br />

are either more competitive in disturbed or moderately polluted soils or<br />

better equipped to survive periods of adverse environmental conditions<br />

than mycorrhizal fungi.<br />

The use of fungicides for crop protection can alter species diversity.<br />

Seed treatment with the systemic fungicide benomyl had no significant<br />

influence on endophyte species richness in wheat roots, but the frequency<br />

of roots colonised by seed borne Septoria nodorum was significantly reduced<br />

(Riesen and Sieber 1985). None of the fungicides applied to Lolium<br />

perenne fields at 18 sites in New Zealand had a significant effect on the<br />

root-endophyte communities (Skipp and Christensen 1989).<br />

Fertilisation can affect fungal assemblages in roots. The frequency of<br />

P. fortinii in seedlings of potted Picea glauca was negatively correlated with<br />

the amount of nitrogen (N) applied (Kernaghan et al. 2003). Wilberforce<br />

et al.(2003) suspected N fertilisers to be one of the mechanisms by which<br />

management affects root endophyte communities in temperate grasslands.<br />

Emissions of air pollutants such as SO2 and especially NOx are thought to<br />

have a similar fertilising effect as fertilisers applied in agriculture. Adverse<br />

effects of these air pollutants on mycorrhizal fungi have been demonstrated<br />

in several studies (Cairney and Meharg 1999; Jansen and van Dobben 1987;<br />

Taylor and Read 1996).


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 119<br />

7.3<br />

Dark Septate Endophytes<br />

Fungi with regularly septate and melanised hyphae probably constitute<br />

the most abundant and most widespread group of non-mycorrhizal root<br />

endophytes.Inthissection,wewillbrieflypresentthehistoryoftheterm<br />

“DSE”, outline the diversity of DSE and give an overview of current knowledge<br />

of the diversity and population genetics of the most prominent species<br />

complex of DSE: Phialocephala fortinii s. l.<br />

7.3.1<br />

History<br />

Melin (1922, 1923) introduced the form taxon Mycelium radicis atrovirens<br />

(MRA) for sterile, melanised, septate mycelia that emerged from mycorrhizae<br />

and roots of Picea abies and Pinus sylvestris. The tree-fungus symbiosis<br />

was characterised by dematiaceous intra- and intercellular hyphae in<br />

the epidermal and cortical cells, but neither a Hartig net nor a mantle were<br />

formed. Melin(1923) coined the term“pseudomycorrhiza” for this relationship<br />

and considered it to form an antagonistic symbiosis. MRA-like fungi<br />

have been detected during numerous studies since Melin’s pioneering work<br />

(Ahlich and Sieber 1996; Chan 1923; Freisleben 1934; Harley and Waid 1955;<br />

Jumpponen et al. 1998; Richard and Fortin 1973; Robertson 1954; Stoyke<br />

and Currah 1991). Since trinomials are not valid species names according to<br />

the International Code of Botanical Nomenclature, less stringent and more<br />

informal names are preferable. Read and Haselwandter (1981) introduced<br />

the term “DS hyphae” (DS = dark septate) for sterile, dark, septate hyphae<br />

and microsclerotia that occurred in roots of various alpine plants. Stoyke<br />

and Currah (1991) implemented the form taxon “dark septate endophyte”<br />

(DSE) and used it for fungi that form partly or entirely melanised, septate<br />

thalli within healthy root tissues. The taxon “DSE” serves primarily to differentiate<br />

these fungi from endophytes with septate, hyaline hyphae, and<br />

from fungi with sparsely septate, hyaline hyphae that are characteristic of<br />

AMF.<br />

7.3.2<br />

Biodiversity<br />

The roots of more than 600 plant species representing about 320 genera in<br />

more than 110 families have been reported to be colonised by DSE (Ahlich<br />

and Sieber 1996; Barrow and Osuna 2002; Jumpponen and Trappe 1998b;


120 T.N. Sieber, C.R. Grünig<br />

Kovacs and Szigetvari 2002; Ruotsalainen et al. 2002; Schadt et al. 2001).<br />

Dematiaceous mycelia are regularly received in culture during censuses of<br />

root endophytes, but it is often not known whether the endophytic thalli of<br />

these fungi are hyaline or melanised. This being the case, we must assume<br />

that DSE are much more widespread than previously assumed.<br />

Species identity of some DSE is known because they readily sporulate<br />

in culture, e.g. Microdochium bolleyi and several Phialophora species in<br />

grasses and sedges. Many non-pathogenic Phialophora endophytes are related<br />

to the take-all fungi (Gaeumannomyces graminis var. tritici and var.<br />

avenae) ofcerealsandgrassesintemperateareasandtoG. graminis var.<br />

graminis, which causes crown sheath rot of rice in the tropics. Phialophora<br />

radicicola forms melanised sclerotia in cortical cells of maize roots without<br />

causing any apparent harm (Cain 1952). P. radicicola was also observed<br />

in the roots of three alpine grasses growing at the timberline in Bavaria<br />

(Blaschke 1986) or in roots of Lolium perenne in New Zealand (Skipp and<br />

Christensen 1989). The DSE abundantly observed in many alpine sedges<br />

in the Tyrolean Alps may also belong to P. radicicola (Haselwandter and<br />

Read 1980; Read and Haselwandter 1981). P. radicicola and P. zeicola,the<br />

maize take-all fungi from China, were recently shown to be the same species<br />

(Ward and Bateman 1999). P. graminicola, another non-pathogenic DSE of<br />

cereal and grass roots (Newsham 1999), provided significant control of the<br />

take-all disease by competition for senescing root tissues (Deacon 1981).<br />

Taxonomic assignment of many DSE is problematic because sexual and<br />

asexual reproductive structures are either absent, rare, or are produced<br />

onlyunderspecificconditions.Coldtreatmentforupto1yearwasshown<br />

to induce sporulation in some DSE isolates, e.g. in isolates of Chloridium<br />

paucisporum, Phialophora finlandica,andPhialocephala fortinii (Wangand<br />

Wilcox 1985). Unfortunately, even then many DSE strains remain sterile<br />

and classification is complicated. Many mycologists have tried to bring<br />

some order into this difficult group of DSE (Harney et al. 1997; Melin<br />

1923; Richard and Fortin 1973). Culture morphology is often used for an<br />

initial classification (Ahlich and Sieber 1996; Girlanda et al. 2002; Steinke<br />

et al. 1996; Stoyke et al. 1992). However, modern molecular biology offers<br />

a multitude of additional and potentially more reliable methods for the<br />

identification and typing of species, varieties and individuals (Carter et al.<br />

1997; Geiser et al. 1994; White et al. 1990; Zietkiewicz et al. 1994). Some of<br />

these methods have been used to type DSE. Restriction patterns of a region<br />

on the ribosomal RNA (rRNA) genes indicated that two-thirds of the DSE<br />

from roots of subalpine plants were closely related to or conspecific with<br />

P. fortinii (Stoyke et al. 1992). Similarly, in a study by Harney et al. (1997),<br />

restriction site mapping of the nuclear rDNA internal transcribed spacer<br />

(ITS) regions showed that the majority of the isolates was P. fortinii-like<br />

and only two isolates were Phialophora finlandica.


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 121<br />

According to isozyme analysis, DSE from various woody plant species<br />

belonged to two distinct groups (Ahlich-Schlegel 1997; Grünig et al. 2001;<br />

Sieber 2002). Members of the larger group were conspecific with P. fortinii,<br />

whereas those of the other group represented the sterile Type 1, which has<br />

been recently described as Acephala applanata (Ahlich and Sieber 1996;<br />

Grünig and Sieber 2005). Phylogenetic analysis of the ITS regions showed<br />

that P. fortinii and A. applanata are closely related and have Phialocephala<br />

compacta, P. dimorphospora and P. scopiformis as closest relatives (Grünig<br />

et al. 2002b). These five species are more closely related to members of the<br />

Leotiales such as Gremmeniella abietina, the causal agent of sclerroderis<br />

canker on pines, than to other Phialocephala species. The “P. fortinii-group”<br />

was also positioned within the Leotiales by phylogenetic analyses of the<br />

sequence data of the 18S and 28S subunits of the nuclear rRNA genes<br />

(Jacobs et al. 2003).<br />

7.3.3<br />

Diversity of Phialocephala fortinii<br />

Phialocephala fortinii was shown to be the dominant DSE in coniferous<br />

and ericaceous roots in heathlands, forests and alpine ecosystems of the<br />

Northern temperate zones (Ahlich and Sieber 1996; Stoyke and Currah<br />

1991). There is strong evidence that the roots of every Norway spruce<br />

(Picea abies) tree in natural forest habitats of Central Europe are colonised<br />

by this fungus (Ahlich and Sieber 1996; Grünig et al. 2004). The nature<br />

of root–P. fortinii symbioses and their ecological significance are largely<br />

unknown.<br />

P. fortinii mayfunctionasamycorrhizalfungusandmediatenutrient<br />

uptake, synthesise secondary metabolites, stimulate plant growth and/or<br />

play an important role in plant defence against root pathogens (Fernando<br />

and Currah 1996; Jumpponen and Trappe 1998a; O’Dell et al. 1993; Yu et al.<br />

2001). Alternatively, it may behave as an opportunistic pathogen (Wilcox<br />

and Wang 1987). However, considering its widespread distribution and<br />

abundanceitisveryunlikelythatP. fortinii is a primary pathogen.<br />

We will provide a compilation of the newest findings on the genetic<br />

diversity within and among populations of P. fortinii and will conclude<br />

this section by forwarding some ideas and thoughts that could explain the<br />

observed diversity of this ecologically very successful species.<br />

Genetic diversity of P. fortinii strains was examined on different spatial<br />

scales using isozymes, PCR-fingerprinting and analysis of the rDNA<br />

ITS regions either by polymerase chain reaction -restriction fragment<br />

length polymorphism (PCR-RFLP) analysis or sequencing. Ahlich-Schlegel<br />

(1997) studied the allelic diversity at seven isozyme loci and detected 108


122 T.N. Sieber, C.R. Grünig<br />

different allozyme phenotypes among 194 European and North-American<br />

DSE strains. Allozyme patterns were neither host- nor site-specific. Harney<br />

et al. (1997) found many polymorphisms in the rDNA ITS regions of<br />

P. fortinii strains from Europe and North America by restriction mapping.<br />

Similarly, variability among rDNA ITS sequences was high (up to 12 substitutions)<br />

among 18 strains of P. fortinii from Central and Northern Europe<br />

(Grünig et al. 2002b). In contrast, Addy et al. (2000) detected a high degree<br />

of homogeneity among the rDNA ITS sequences of six strains of P. fortinii<br />

from Canada and Japan.<br />

Strain-specific markers are necessary to study the genetic diversity at<br />

small spatial scales. In contrast to allozyme markers, ISSR-PCR markers<br />

were strain specific and allowed discrimination among isolates with identical<br />

allozyme phenotypes (Grünig et al. 2001). These markers were used to<br />

detect the population structure of DSE isolated from Norway spruce (Picea<br />

abies) roots collected within a 3 × 3 m plot of a 40-year-old plantation<br />

(Grünig et al. 2002a). Twenty-one unique ISSR-PCR genets were present<br />

among 144 strains. Identity of the isolated DSE as P. fortinii was confirmed<br />

by the morphology of the conidiogenous apparatus and by sequence comparisons<br />

of the rDNA ITS regions. Two genets dominated and were isolated<br />

from all sampling points within contiguous areas of at least 6.8 m 2 and<br />

5.3 m 2 that overlapped by 3.6 m 2 .Othergenetswererareandwereisolated<br />

only once or twice.<br />

Jumpponen (1999) employed the random amplified polymorphic DNA<br />

(RAPD) technique to determine the population structure of P. fortinii at<br />

a primary succession site on a glacier forefront. In one year, 23 genets of<br />

P. fortinii were detected in 34 strains, in the next year 10 genets were found<br />

in 40 strains, but none of the genets was isolated in both years. Diversity<br />

of P. fortinii canbehighevenwithinsinglerootpieces.Forexample,8to<br />

10-cm-long pieces of fine root of Picea abies were colonised by up to<br />

six different inter-simple sequence repeat (ISSR) phenotypes (N. Nüssli<br />

and C.R. Grünig, unpublished) (Fig. 7.1). In summary, genetic diversity of<br />

P. fortinii seems to be high at every level. This is surprising for a supposedly<br />

asexual fungus. Therefore, studies on population genetics were initiated to<br />

findthesourcesofthishighdiversity.<br />

ISSR-PCR and RAPD markers have many analytical drawbacks, such<br />

as dominance, and they cannot be used to infer population differentiation<br />

and recombination. In contrast, single-locus RFLP markers are codominant<br />

and supply robust data for precise population genetic analyses. In addition,<br />

data are comparable among studies and thus may be used for global analyses<br />

(Sunnucks 2000). Therefore, single-locus RFLP probes were developed<br />

for population genetic analysis of P. fortinii and used to find evidence for<br />

recombination, gene and genotype flow in P. fortinii (Grünig et al. 2003,<br />

2004). Strains collected from three Norway-spruce plots up to 10 km apart


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 123<br />

Fig.7.1. Distribution of six inter-simple sequence repeat-polymerase chain reaction (ISSR-<br />

PCR) phenotypes belonging to three cryptic species of the root endophyte Phialocephala<br />

fortinii s. l. along a healthy fine root of Norway spruce (Picea abies). Identical symbols<br />

indicate positions on the root where the same phenotype was isolated. Symbols with identical<br />

shape represent the same cryptic species. C Positions on the root where Cylindrocarpon<br />

didymum wasisolatedasanendophyte<br />

from each other were studied using 11 single-locus RFLP probes. The average<br />

gene diversity was high and up to 96 multilocus haplotypes (MLH) were<br />

observed per study plot. Significant population subdivision was detected<br />

among groups of MLH within plots, suggesting that groups were reproductively<br />

isolated and should be considered cryptic species. The RFLP data of<br />

more than 1,000 European strains indicate that P. fortinii s. l. is a species<br />

complex of at least eight cryptic species (C.R. Grünig, unpublished). The<br />

index of association (IA) did not deviate significantly from zero within any<br />

cryptic species, suggesting that recombination occurs, or has occurred,<br />

within these species. Although evidence for recombination is strong for all<br />

cryptic species, it remains unclear whether sexual or parasexual processes<br />

are involved, and how often and where recombination occurs or when it last<br />

occurred (Taylor et al. 1999). Even a little sex is, however, already enough<br />

to give an organism the appearance of a recombining population (Brown<br />

1999).<br />

The sympatric occurrence of up to four reproductively isolated, cryptic<br />

species within a few square metres of forest floor, and sometimes even<br />

in the same root segment, is a highly interesting phenomenon and deserves<br />

a brief discussion (Grünig et al. 2004) (Figs. 7.1, 7.2). Reproductive<br />

isolation is essential for speciation. Geographically isolated populations<br />

are often reproductively isolated, and may experience allopatric speciation<br />

through genetic drift (Carter et al. 2001). On the other hand, niche or<br />

habitat specialisation may lead to sympatric speciation when local populations<br />

are confronted with heterogeneous habitats or several niches within<br />

habitats (Futuyma and Moreno 1988; Maynard Smith 1966). The patterns<br />

observed by Grünig et al.(2004) are clearly indicative of speciation. Possibly,<br />

the cryptic species were the products of allopatric speciation in the


124 T.N. Sieber, C.R. Grünig<br />

Fig.7.2. Distribution of the four most frequently observed cryptic species (csp) ofPhialocephala<br />

fortinii s. l. within healthy fine roots of Norway spruce (Picea abies) collectedat<br />

the intersections of a 2 × 2 m grid superimposed on a forest plot (14 × 14 m) at Zürichberg,<br />

Switzerland. The four graphs represent the same study plot; the distributions of the four<br />

cryptic species are presented in separate graphs to maintain clarity. The number of isolates<br />

and the multilocus haplotypes (MLH) of each cryptic species are given in brackets<br />

past due to geographical isolation. The ranges of these species may have<br />

subsequently overlapped (Brasier 1987). In this respect it is interesting to<br />

study the role of Quaternary climatic changes (Hewitt 2000). The succession<br />

of several glaciations and warmer inter-glacial periods had profound<br />

effects on animals, plants, and, consequently, on fungi. During the Quaternary,<br />

each species experienced many contractions/expansions of range,<br />

leading to extinctions and foundations of populations, decreases and increases<br />

in diversity and, thus, also to speciation (Taberlet et al. 1998).<br />

Refugia of relevant hosts of P. fortinii were often geographically isolated,


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 125<br />

making allopatric speciation of P. fortinii possible. Alternatively, habitat<br />

heterogeneities are certainly present even within very small compartments<br />

of root tissues, the rhizosphere, and the surrounding soil. These heterogeneities<br />

may be pronounced enough for ecological isolation and for the<br />

development of cryptic species. Some cryptic species may be interspecific<br />

hybrids. For example, most asexual Epichloë-related grass endophytes appear<br />

to be such hybrids (Scott 2001). Interspecific hybrids may be better<br />

adapted to new niches such as new hosts and can provide greater or more<br />

diverse benefits to host plants (Schardl and Craven 2003). However, such<br />

hybrids were never observed for P. fortinii using codominantly inherited<br />

single-copy RFLP markers.<br />

MLHwithidenticalISSRfingerprintingpatternswerecommontoatleast<br />

two of the sites in the study of Grünig et al. (2004). These results indicate<br />

that not only gene flow but also genotype flow most likely occurs in cryptic<br />

species of P. fortinii. Gene and genotype flow occur either naturally via<br />

conidia or microsclerotia transported by wind or micro- and macrofauna,<br />

or by silvicultural practices. Genotypes may be introduced by planting<br />

plants from nurseries located up to several hundreds of kilometers away<br />

(Bürgi and Schuler 2003), since nursery plants are frequently colonised<br />

by DSE including P. fortinii (Danielson and Visser 1990). Alternatively,<br />

machinery used during thinning and harvesting could be responsible for<br />

the import of genotypes.<br />

Nothing is known about the significance of mutations, the ultimate<br />

source of genetic variation, for speciation within P. fortinii s. l. If a population<br />

is large and the mutation rate high, it is likely that mutants with<br />

higher fitness, e.g. better mutualists, will emerge (McDonald and Linde<br />

2002). Non-lethal somatic mutations in the mitotic phase may affect the genetic<br />

diversity of a population since each nucleus has the capacity to be the<br />

founder genome of another, new mycelium (Burnett 2003). The diversity<br />

thus generated may supplement diversity generated by recombination.<br />

7.4<br />

Conclusions<br />

Colonisation of roots by fungal endophytes is a common feature in the plant<br />

kingdom. In contrast to classical mycorrhizae, endophytes are regularly<br />

present in roots undergoing secondary growth. Root-endophyte species<br />

diversity is affected by climatic, physical, chemical, biological and anthropogenic<br />

factors. DSEs are among the most abundant root endophytes.<br />

They constitute a taxonomically very heterogeneous group of fungi, mostly<br />

ascomycetes, that form melanised, septate hyphae, chlamydospores or microsclerotia<br />

within the roots of the host.


126 T.N. Sieber, C.R. Grünig<br />

Phialocephala fortinii is the most prominent DSE, especially in woody<br />

plant species. P. fortinii s.l.isgenotypicallyverydiverseandformsacomplex<br />

of several cryptic species that can occur sympatrically. Cryptic species<br />

and selected genotypes of P. fortinii s. l. can now be used to test the ecological<br />

significance of these extremely abundant and successful organisms and<br />

to explain some of the contradictory results on fungus-host interactions<br />

reported in earlier studies. The elucidation of the mating mechanism(s)<br />

and the evolutionary forces that govern speciation in P. fortinii s. l. are<br />

other fascinating topics for future research.<br />

We have reviewed patterns of species diversity and within-species genotypic<br />

diversity and presented several plausible explanations for these patterns,<br />

although conclusive evidence for cause and effect are still virtually<br />

lacking. Nevertheless, we would like to conclude with a motivating citation<br />

by Begon et al. (1990): “This is not so much a disappointment as a challenge<br />

to ecologists and biologists of the future. Much of the fascination of ecology<br />

and biology lies in the fact that many problems are blatant and obvious for<br />

everybody to see, while the solutions have as yet eluded us”.<br />

<strong>References</strong><br />

Addy HD, Hambleton S, Currah RS (2000) Distribution and molecular characterization of<br />

the root endophyte Phialocephala fortinii along an environmental gradient in the boreal<br />

forest of Alberta. Mycol Res 104:1213–1221<br />

Ahlich K, Sieber TN (1996) The profusion of dark septate endophytic fungi in nonectomycorrhizal<br />

fine roots of forest trees and shrubs. New Phytol 132:259–270<br />

Ahlich K, Rigling D, Holdenrieder O, Sieber TN (1998) Dark septate hyphomycetes in Swiss<br />

conifer forest soils surveyed using Norway-spruce seedlings as bait. Soil Biol Biochem<br />

30:1069–1075<br />

Ahlich-Schlegel K (1997) Vorkommen und Charakterisierung von dunklen, septierten Hyphomyceten<br />

(DSH) in Gehölzwurzeln. PhD dissertation, Swiss Federal Institute of Technology,<br />

Department of Forest Sciences, Zürich, Switzerland<br />

Ananda K, Sridhar KR (2002) Diversity of endophytic fungi in the roots of mangrove species<br />

on the west coast of India. Can J Microbiol 48:871–878<br />

Bååth E, Söderström B (1979) Fungal biomass and fungal immobilization of plant nutrients<br />

in Swedish coniferous forest soils. Rev Ecol Biol Sol 16:477–489<br />

Barrow JR, Osuna P (2002) Phosphorus solubilization and uptake by dark septate fungi in<br />

fourwing saltbush, Atriplex canescens (Pursh) Nutt. J Arid Environ 51:449–459<br />

Bazin MJ, Markham P, Scott EM, Lynch JM (1990) Population dynamics and rhizosphere<br />

interactions. In: Lynch JM (ed) The rhizosphere. Wiley, Chichester, UK, pp 99–127<br />

Begon M, Harper JL, Townsend CR (1990) Ecology – individuals, populations, communities,<br />

2ndedn.Blackwell,Oxford,UK<br />

Bever JD, Westover KM, Antonovics J (1997) Incorporating the soil community into plant<br />

population dynamics: the utility of the feedback approach. J Ecol 85:561–573<br />

Blaschke H (1986) Vergleichende Untersuchungen über die Entwicklung mykorrhizierter<br />

Feinwurzeln von Fichten in Waldschadensgebieten. Forstw Cbl 105:477–487<br />

Bledsoe C, Klein P, Bliss LC (1990) A survey of mycorrhizal plants on Truelove Lowland,<br />

Devon Island, N. W. T., Canada. Can J Bot 68:1848–1856


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 127<br />

Brasier CM (1987) The dynamics in fungal speciation. In: Rayner ADM, Brasier CM, Moore D<br />

(eds) Evolutionary biology of the fungi. Cambridge University Press, Cambridge, UK,<br />

pp 232–260<br />

Bridge P, Spooner B (2001) Soil fungi: diversity and detection. Plant Soil 232:147–154<br />

Brown JKM (1999) The evolution of sex and recombination in fungi. In: Worrall JJ (ed)<br />

Structure and dynamics of fungal populations. Kluwer, Dordrecht, pp 73–95<br />

Bruns TD, Bidartondo MI, Taylor DL (2002) Host specificity in ectomycorrhizal communities:<br />

What do the exceptions tell us? Integr Comp Biol 42:352–359<br />

Bürgi M, Schuler A (2003) Driving forces of forest management – an analysis of regeneration<br />

practices in the forests of the Swiss Central Plateau during the 19th and 20th century.<br />

For Ecol Manage 176:173–183<br />

Burnett J (2003) Fungal populations and species. Oxford University Press, Oxford, UK<br />

Cain RF (1952) Studies of fungi imperfecti. I. Phialophora. Can J Bot 30:338–343<br />

Cairney JWG, Meharg AA (1999) Influences of anthropogenic pollution on mycorrhizal<br />

fungal communities. Environ Pollut 106:169–182<br />

Cannon PF (1997) Strategies for rapid assessment of fungal diversity. Biodivers Conserv<br />

6:669–680<br />

Cannon PF, Hawksworth DL (1995) The diversity of fungi associated with vascular plants:<br />

the known, the unknown and the need to bridge the knowledge gap. Adv Plant Pathol<br />

11:277–302<br />

Carroll GC (1988) Fungal endophytes in stems and leaves – from latent pathogen to mutualistic<br />

symbiont. Ecology 69:2–9<br />

Carter DA, Burt A, Taylor JW, Koenig GL, Dechairo B, White TJ (1997) A set of electrophoretic<br />

molecular markers for strain typing and population genetic studies of Histoplasma<br />

capsulatum. Electrophoresis 18:1047–1053<br />

Carter DA, Taylor JW, Dechairo B, Burt S, Koenig GL, White TJ (2001) Amplified singlenucleotide<br />

polymorphisms and a (GA)(n) microsatellite marker reveal genetic differentiation<br />

between populations of Histoplasma capsulatum from the Americas. Fungal<br />

Genet Biol 34:37–48<br />

Cevnik M, Jurc M, Vodnik D (2000) Filamentous fungi associated with the fine roots of Erica<br />

herbacea L. from the area influenced by the Zerjav lead smelter (Slovenia). Phyton Ann<br />

Rei Bot 40:61–64<br />

Chan TAB (1923) Über die Mykorrhiza der Buche. Allg Forst J Ztg 99:25–52<br />

Christensen M (1981) Species diversity and dominance in fungal communities. In: Wicklow<br />

DT, Carroll GC (eds) The fungal community, its organization and role in the ecosystem.<br />

Dekker, New York, pp 201–232<br />

Christensen M (1989) A view of fungal ecology. Mycologia 81:1–19<br />

Christie P, Nicolson TH (1983) Are mycorrhizas absent from the Antarctic? Trans Br Mycol<br />

Soc 80:557–560<br />

Coûteaux M-M, Kurz C, Bottner P, Raschi A (1999) Influence of increased atmospheric<br />

CO2 concentration on quality of plant material and litter decomposition. Tree Physiol<br />

19:301–311<br />

Danielson RM, Visser S (1990) The mycorrhizal and nodulation status of container-grown<br />

trees and shrubs reared in commercial nurseries. Can J For Res 20:609–614<br />

Deacon JW (1981) Ecological relationships with other fungi: competitors and hyperparasites.<br />

In: Asher MJC, Shipton PJ (eds) Biology and control of Take-all. Academic, London,<br />

UK, pp 75–101<br />

Fernando AA, Currah RS (1996) A comparative study of the effects of the root endophytes<br />

Leptodontidium orchidicola and Phialocephala fortinii (fungi imperfecti) on the growth<br />

of some subalpine plants in culture. Can J Bot 74:1071–1078


128 T.N. Sieber, C.R. Grünig<br />

Fisher PJ, Petrini O, Webster J (1991) Aquatic hyphomycetes and other fungi in living aquatic<br />

and terrestrial roots of Alnus glutinosa. Mycol Res 95:543–547<br />

Fisher PJ, Graf F, Petrini LE, Sutton BC, Wookey PA (1995) Fungal endophytes of Dryas<br />

octopetala from a high arctic polar semidesert and from the Swiss Alps. Mycologia<br />

87:319–323<br />

Fitter AH, Atkinson D, Read DJ, Usher MB (1985) Ecological interactions in soil. Blackwell,<br />

Oxford<br />

Foster RC, Rovira AD, Cock TW (1983) Ultrastructure of the root-soil interface. APS Press,<br />

St. Paul, MN<br />

Freisleben R (1934) Zur Frage der Mykotrophie in der Gattung Vaccinium L. Jahrb wissenschaftl<br />

Bot 80:421–456<br />

Futuyma DJ, Moreno G (1988) The evolution of ecological specialization. Annu Rev Ecol<br />

Syst 19:207–233<br />

Geiser DM, Arnold ML, Timberlake WE (1994) Sexual origins of British Aspergillus nidulans<br />

isolates. Proc Natl Acad Sci USA 91:2349–2352<br />

Girlanda M, Ghignone S, Luppi AM (2002) Diversity of sterile root-associated fungi of two<br />

Mediterranean plants. New Phytol 155:481–498<br />

Gochenaur SE (1984) Fungi of a Long Island oak-birch forest. II. Population dynamics and<br />

hydrolase patterns for the soil Penicillia. Mycologia 76:218–231<br />

Görke C (1998) Mykozönosen von Wurzeln und Stamm von Jungbäumen unterschiedlicher<br />

Bestandsbegründungen. Bibl Mycol 173:1–462<br />

Grünig CR, Sieber TN (2005) Molecular and phenotypic description of the widespread<br />

root symbiont Acephala applanata gen. et sp. nov., formerly known as “Dark Septate<br />

Endophyte Type 1”. Mycologia 97:628–640<br />

Grünig CR, Sieber TN, Holdenrieder O (2001) Characterisation of dark septate endophytic<br />

fungi (DSE) using inter-simple-sequence-repeat-anchored polymerase chain reaction<br />

(ISSR-PCR) amplification. Mycol Res 105:24–32<br />

Grünig CR, Sieber TN, Rogers SO, Holdenrieder O (2002a) Spatial distribution of dark<br />

septate endophytes in a confined forest plot. Mycol Res 106:832–840<br />

Grünig CR, Sieber TN, Rogers SO, Holdenrieder O (2002b) Genetic variability among strains<br />

of Phialocephala fortinii and phylogenetic analysis of the genus Phialocephala based on<br />

rDNA ITS sequence comparisons. Can J Bot 80:1239–1249<br />

Grünig CR, Linde CC, Sieber TN, Rogers SO (2003) Development of single-copy RFLP<br />

markers for population genetic studies of Phialocephala fortinii and closely related<br />

taxa. Mycol Res 107:1332–1341<br />

Grünig CR, McDonald BA, Sieber TN, Rogers SO, Holdenrieder O (2004) Evidence for<br />

subdivision of the root-endophyte Phialocephala fortinii into cryptic species and recombination<br />

within species. Fungal Genet Biol 41:676–687<br />

Hadacek F, Kraus GF (2002) Plant root carbohydrates affect growth behaviour of endophytic<br />

microfungi. FEMS Microbiol Ecol 41:161–170<br />

Harley JL, Waid JS (1955) A method of studying active mycelia on living roots and other<br />

surfaces in soil. Trans Br Mycol Soc 38:104–118<br />

Harney SK, Rogers SO, Wang CJK (1997) Molecular characterization of dematiaceous root<br />

endophytes. Mycol Res 101:1397–1404<br />

Haselwandter K, Read DJ (1980) Fungal associations of roots of dominant and sub-dominant<br />

plants in high-alpine vegetation systems with special reference to mycorrhiza. Oecologia<br />

45:57–62<br />

Haselwandter K, Read DJ (1982) The significance of a root-fungus association in two Carex<br />

species of high-alpine plant communities. Oecologia 53:352–354


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 129<br />

Hashimoto Y, Hyakumachi M (2000) Quantities and types of ectomycorrhizal and endophytic<br />

fungi associated with Betula platyphylla var. japonica seedlings during the initial<br />

stage of establishment of vegetation after disturbance. Ecol Res 15:21–31<br />

Hawksworth DL (1991) The fungal dimension of biodiversity: magnitude, significance, and<br />

conservation. Mycol Res 95:641–655<br />

Hewitt GM (2000) The genetic legacy of the Quaternary ice ages. Nature 405:907–913<br />

Hill MO (1973) Diversity and evenness: a unifying notation and its consequences. Ecology<br />

54:427–432<br />

Horton TR, Cázares E, Bruns TD (1998) Ectomycorrhizal, vesicular-arbuscular and dark septate<br />

fungal colonisation of bishop pine (Pinus muricata) seedlings in the first 5 months<br />

of growth after wildfire. Mycorrhiza 8:11–18<br />

Hurlbert SH (1971) The non-concept of species diversity: a critique and alternative parameters.<br />

Ecology 52:577–586<br />

Jacobs A, Coetzee MPA, Wingfield BD, Jacobs K, Wingfield MJ (2003) Phylogenetic relationships<br />

among Phialocephala species and other ascomycetes. Mycologia 95:637–645<br />

Jansen E, van Dobben HF (1987) Is decline of Cantharellus cibarius in the Netherlands due<br />

to air pollution. Ambio 16:27–29<br />

Jones MD, Durall DM, Harniman SMK, Classen DC, Simard SW (1997) Ectomycorrhizal<br />

diversity on Betula papyrifera and Pseudotsuga menziesii seedlings grown in the greenhouse<br />

or outplanted in single-species and mixed plots in southern British Columbia.<br />

Can J For Res 27:1872–1889<br />

Jumpponen A (1999) Spatial distribution of discrete RAPD phenotypes of a root endophytic<br />

fungus, Phialocephala fortinii, at a primary successional site on a glacier forefront. New<br />

Phytol 141:333–344<br />

Jumpponen A, Trappe JM (1998a) Performance of Pinus contorta inoculated with two strains<br />

of root endophytic fungus, Phialocephala fortinii: effects of synthesis system and glucose<br />

concentration. Can J Bot 76:1205–1213<br />

Jumpponen A, Trappe JM (1998b) Dark septate endophytes: A review of facultative<br />

biotrophic root-colonising fungi. New Phytol 140:295–310<br />

Jumpponen A, Mattson KG, Trappe JM (1998) Mycorrhizal functioning of Phialocephala<br />

fortinii with Pinus contorta on glacier forefront soil: interactions with soil nitrogen and<br />

organic matter. Mycorrhiza 7:261–265<br />

Jumpponen A, Claridge AW, Trappe JM, Lebel T, Claridge DL (2004) Ecological relationships<br />

among hypogeous fungi and trees: inferences from association analysis integrated with<br />

habitat modeling. Mycologia 96:510–525<br />

Kattner D, Schönhar S (1990) Untersuchungen über das Vorkommen mikroskopischer Pilze<br />

in Feinwurzeln optisch gesunder Fichten (Picea abies Karst.) auf verschiedenen Standorten.<br />

Mitt Ver Forstl Standortskde Forstpflanzenzücht 35:39–43<br />

Kernaghan G, Sigler L, Khasa D (2003) Mycorrhizal and root endophytic fungi of containerized<br />

Picea glauca seedlings assessed by rDNA sequence analysis. Microb Ecol<br />

45:128–136<br />

Kirk JJ, Deacon JW (1987) Control of the take-all fungus by Microdochium bolleyi, and<br />

interactions involving M. bolleyi, Phialophora graminicola and Periconia macrospinosa<br />

on cereal roots. Plant Soil 98:231–237<br />

Körner C (2000) Biosphere responses to CO2 enrichment. Ecol Appl 10:1590–1619<br />

Kovacs GM, Szigetvari C (2002) Mycorrhizae and other root-associated fungal structures<br />

of the plants of a sandy grassland on the Great Hungarian Plain. Phyton Ann Rei Bot<br />

42:211–223<br />

Maynard Smith J (1966) Sympatric speciation. Am Nat 100:637–650


130 T.N. Sieber, C.R. Grünig<br />

McDonald BA, Linde C (2002) Pathogen population genetics, evolutionary potential, and<br />

durable resistance. Annu Rev Phytopathol 40:349–379<br />

Melin E (1922) On the mycorrhizas of Pinus silvestris L. and Picea abies (L.) Karst. J Ecol<br />

9:254–257<br />

Melin E (1923) Experimentelle Untersuchungen über die Konstitution und Ökologie der<br />

Mycorrhizen von Pinus silvestris L. und Picea abies (L.) Karst. Falk Mykol Unters<br />

2:73–331<br />

Narisawa K, Kawamata H, Currah RS, Hashiba T (2002) Suppression of Verticillium wilt in<br />

eggplant by some fungal root endophytes. Eur J Plant Pathol 108:103–109<br />

Newsham KK (1999) Phialophora graminicola, a dark septate fungus, is a beneficial associate<br />

of the grass Vulpia ciliata ssp. ambigua. New Phytol 144:517–524<br />

Oba H, Tawaraya K, Wagatsuma T (2002) Inhibition of pre-symbiotic hyphal growth of<br />

arbuscular mycorrhizal fungus Gigaspora margarita by root exudates of Lupinus spp.<br />

Soil Sci Plant Nutr 48:117–120<br />

Oberholzer-Tschütscher B (1982) Untersuchungen über endophytische Pilze von Erica<br />

carnea L. PhD dissertation, Swiss Federal Institute of Technology, Institute of Microbiology,<br />

Zürich, Switzerland<br />

O’Dell TE, Massicotte HB, Trappe JM (1993) Root colonisation of Lupinus latifolius Agardh.<br />

and Pinus contorta Dougl. by Phialocephala fortinii Wang&Wilcox.NewPhytol<br />

124:93–100<br />

Papritz A, Flühler H (1991) Räumliche Verteilung von bodenchemischen Grössen auf<br />

Transsekten zwischen Bäumen (Beobachtungsfläche Lägern). In: Pankow W (ed)<br />

Lufthaushalt, Luftverschmutzung und Waldschäden in der Schweiz, Band 6, Belastung<br />

von Waldböden. Verlag der Fachvereine, Zürich, pp 125–136<br />

Petrini O, Fisher PJ, Petrini LE (1992) Fungal endophytes of bracken (Pteridium aquilinum),<br />

with some reflections on their use in biological control. Sydowia 44:282–293<br />

Read DJ, Haselwandter K (1981) Observation on the mycorrhizal status of some alpine plant<br />

communities. New Phytol 88:341–352<br />

Reinecke P (1978) Microdochium bolleyi at the stem base of cereals. Z Pflanzenkr Pflanzenschutz<br />

85:679–685<br />

Richard C, Fortin J-A (1973) The identification of Myceliumradicis atrovirens (Phialocephala<br />

dimorphospora). Can J Bot 51:2247–2248<br />

Riesen T, Sieber TN(1985) Endophyticfungi in winter wheat(Triticum aestivum L.). Institute<br />

of Microbiology, Swiss Federal Institute of Technology, Zürich, Switzerland<br />

Rillig MC, Wright SF, Allen MF, Field CB (1999) Rise in carbon dioxide changes in soil<br />

structure. Nature 400:628–628<br />

Robertson NF (1954) Studies on the mycorrhiza of Pinus silvestris.I.Patternofdevelopment<br />

of mycorrhizal root and its significance for experimental studies. New Phytol 53:253–283<br />

Ruotsalainen AL, Väre H, Vestberg M (2002) Seasonality of root fungal colonisation in<br />

low-alpine herbs. Mycorrhiza 12:29–36<br />

Schadt CW, Mullen RB, Schmidt SK (2001) Isolation and phylogenetic identification of<br />

a dark-septate fungus associated with the alpine plant Ranunculus adoneus.NewPhytol<br />

150:747–755<br />

Schardl CL, Craven KD (2003) Interspecific hybridization in plant-associated fungi and<br />

oomycetes: a review. Mol Ecol 12:2861–2873<br />

Schulz B, Römmert AK, Dammann U, Aust H-J, Strack D (1999) The endophyte-host interaction:<br />

a balanced antagonism? Mycol Res 103:1275–1283<br />

Scott B (2001) Epichloë endophytes: fungal symbionts of grasses. Curr Opin Microbiol<br />

4:393–398<br />

Sieber TN (2002) Fungal root endophytes. In: Waisel Y, Eshel A, Kafkafi U (eds) Plant roots:<br />

The hidden half, 3rd edn. Dekker, New York, pp 887–917


7 Biodiversity of Fungal Root-Endophyte Communities and Populations 131<br />

Sieber TN, Riesen TK, Müller E, Fried PM (1988) Endophytic fungi in four winter wheat<br />

cultivars (Triticum aestivum L.) differing in resistance against Stagonospora nodorum<br />

(Berk.) Cast. & Germ. = Septoria nodorum (Berk.) Berk. J Phytopathol 122:289–306<br />

Skipp RA, Christensen MJ (1989) Fungi invading roots of perennial ryegrass (Lolium<br />

perenne L.) in pasture. N Z J Agric Res 32:423–431<br />

Soerensen T (1948) A method of establishing groups of equal amplitude in plant sociology.<br />

Vid Selsk Biol Skr 5:4–4<br />

Steinke E, Williams PG, Ashford AE (1996) The structure and fungal associates of mycorrhizas<br />

in Leucopogon parviflorus (Andr.) Lindl. Ann Bot 77:413–419<br />

Stoyke G, Currah RS (1991) Endophytic fungi from the mycorrhizae of alpine ericoid plants.<br />

Can J Bot 69:347–352<br />

Stoyke G, Egger KN, Currah RS (1992) Characterization of sterile endophytic fungi from<br />

the mycorrhizae of subalpine plants. Can J Bot 70:2009–2016<br />

Sunnucks P (2000) Efficient genetic markers for population biology. Trends Ecol Evol<br />

15:199–203<br />

Swift MJ (1976) Species diversity and the structure of microbial communities in terrestrial<br />

habitats. In: Anderson JM, Macfayden A (eds) The role of terrestrial and aquatic<br />

organisms in decomposition processes. Blackwell, Oxford, UK, pp 185–222<br />

Taberlet P, Fumagalli L, Wust-Saucy AG, Cosson JF (1998) Comparative phylogeography<br />

and postglacial colonisation routes in Europe. Mol Ecol 7:453–464<br />

Taylor AFS, Read DJ (1996) A European north-south survey of ectomycorrhizal populations<br />

on spruce. In: Azcon-Aguilar C, Barea JM (eds) Mycorrhizas in integrated systems from<br />

genes to plant development, Proc 4th European Symposium on Mycorrhizas. Office for<br />

official publications of the European Community, Luxembourg, pp 144–147<br />

Taylor JW, Jacobson DJ, Fisher MC (1999) The evolution of asexual fungi: reproduction,<br />

speciation and classification. Annu Rev Phytopathol 37:197–246<br />

Torsvik V, Salte K, Sorheim R, Goksoyr J (1990) Comparison of phenotypic diversity and<br />

DNA heterogeneity in a population of soil bacteria. Appl Environ Microbiol 56:776–781<br />

Van Der Putten WH (2003) Plant defense belowground and spatiotemporal processes in<br />

natural vegetation. Ecology 84:2269–2280<br />

VanTolRWHM,VanderSommenATC,BoffMIC,VanBezooijenJ,SabelisMW,SmitsPH<br />

(2001) Plants protect their roots by alerting the enemies of grubs. Ecol Lett 4:292–294<br />

Väre H, Vestberg M, Eurola S (1992) Mycorrhiza and root associated fungi in Spitsbergen.<br />

Mycorrhiza 1:93–104<br />

Wang CJK, Wilcox HE (1985) New species of ectendomycorrhizal and pseudomycorrhizal<br />

fungi: Phialophora finlandia, Chloridium paucisporum,andPhialocephala fortinii.Mycologia<br />

77:951–958<br />

Ward E, Bateman GL (1999) Comparison of Gaeumannomyces-andPhialophora-like fungal<br />

pathogens from maize and other plants using DNA methods. New Phytol 141:323–331<br />

Watanabe T (1994) Pictorial atlas of soil and seed fungi. Lewis, Boca Raton, FL<br />

White TJ, Bruns TD, Lee S, Taylor J (1990) Amplification and direct sequencing of fungal<br />

ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ<br />

(eds) PCR Protocols: a guide to methods and applications. Academic, New York, pp 315–<br />

322<br />

Wilberforce EM, Boddy L, Griffiths R, Griffith GW (2003) Agricultural management affects<br />

communities of culturable root-endophytic fungi in temperate grasslands. Soil Biol<br />

Biochem 35:1143–1154<br />

Wilcox HE, Wang CJK (1987) Mycorrhizal and pathological associations of dematiaceous<br />

fungi in roots of 7-month-old tree seedlings. Can J For Res 17:884–899<br />

Wirsel SGR, Leibinger W, Ernst M, Mendgen K (2001) Genetic diversity of fungi closely<br />

associated with common reed. New Phytol 149:589–598


132 T.N. Sieber, C.R. Grünig<br />

Yu T, Nassuth A, Peterson RL (2001) Characterization of the interaction between the dark<br />

septate fungus Phialocephala fortinii and Asparagus officinalis roots. Can J Microbiol<br />

47:741–753<br />

Zietkiewicz E, Rafalski A, Labuda D (1994) Genome fingerprinting by simple sequence<br />

repeat (SSR)-anchored polymerase chain reaction amplification. Genomics 20:176–183


8<br />

Endophytic Root Colonization<br />

by Fusarium Species: Histology,<br />

Plant Interactions, and Toxicity<br />

Charles W. Bacon, Ida E. Yates<br />

8.1<br />

Introduction<br />

Fusarium species have adapted to a wide range of geographical sites, climatic<br />

conditions, ecological habitats, and host plants, and species of this<br />

polyphyletic genus have been documented to occur worldwide (Backhouse<br />

et al. 2001). In spite of the information available on the extremes in geographic<br />

distribution and climatic conditions, appropriate data to predict<br />

the center of origin(s) or the mode(s) of dispersion of this genus have<br />

not been obtained. Much of the information on distribution patterns has<br />

been determined from analyses of soil samples, a common habitat, in addition<br />

to colonization of many plant species. The diversity of plant species<br />

colonized by members of the genus Fusarium is amazing. A recent literature<br />

survey determined that Fusarium species have been isolated from<br />

plants belonging to the gymnosperms and the monocotyledonous and<br />

dicotyledonous angiosperms (Kuldau and Yates 2000). They are the primary<br />

incitants of root, stem, and ear rots in many agriculturally important<br />

crops. For example, F. verticillioides (= F. moniliforme) is capable of colonizing<br />

well over 1,000 plant species, including maize (Zea mays L.), one<br />

of the world’s most important food crops. Another species, F. oxysporum,<br />

is cosmopolitan; certain strains are usually host specific and pose a severe<br />

threat to most of the world’s supply of food crops. Furthermore, species<br />

such as F. graminearum, along with F. verticillioides and related species<br />

within the Liseola section, are notorious for the production of mycotoxins<br />

on wheat, maize, barley, rice and other cereal grains and foodstuffs<br />

(Marasas et al. 1984). Consequently, studies on the association of Fusarium<br />

species with plants are critical in order to develop control measures for<br />

this group of fungi that affects the quality and quantity of the world’s food<br />

supply.<br />

Charles W. Bacon: Richard B. Russell Research Center, ARS, United States Department of<br />

Agriculture, Toxicology and Mycotoxin Research Unit, SAA, P.O. Box 5677, Athens, GA<br />

30604, USA, E-mail: cbacon@saa.ars.usda.gov<br />

Ida E. Yates: Richard B. Russell Research Center, ARS, United States Department of Agriculture,<br />

Toxicology and Mycotoxin Research Unit, Athens, GA 30604, USA<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


134 C.W. Bacon, I.E. Yates<br />

The discussion of Fusarium root endophytes in this chapter is based<br />

on, and will be discussed relative to, our knowledge of fungal endophytes<br />

of grasses. Noted examples of fungal endophytes include the species of<br />

the Balansieae that show various degrees of tissue specificity and often<br />

display evidence of infection by the production of sporulation structures<br />

on the adaxial or abaxial leaf surface of grasses (Diehl 1950), as well as the<br />

productionofcharacteristictoxicsecondary metabolites(Baconetal. 1986).<br />

For example, fungi of the genus Neotyphodium (teleomorph = Epichloë)<br />

arefoundonlyinthestems,leavesandseedofgrassesbutarenotfound<br />

in roots, while species of Myriogenospora are restricted to the leaves, and<br />

species of Balansia are restricted to stems or leaves, but all produce ergot<br />

alkaloids.<br />

Some research suggests that there are similar positive interactions of<br />

endophytic Fusarium species with plants (Damicone and Manning 1982;<br />

Hallmann and Sikora 1994a, 1994b; Blok and Bollen 1995). We use the<br />

definition of fungal endophytes as indicated by Stone et al. (2000) to include<br />

those Fusarium species that are associated with roots as intercellular,<br />

symptomless fungi (Fig. 8.1a–e), although the endophytic association may<br />

extend to above ground plant parts and there may be a differential expression<br />

of infection with different host tissue types (Bergman and Bakker-Van<br />

der Voort 1979; Fisher et al. 1992; Foley 1962; Yates and Jaworski 2000).<br />

In addition to grass endophytes, specific attention will be directed to our<br />

past and present toxicological, physiological, and morphological studies on<br />

Fusarium verticillioides [synonym F. moniliforme, teleomorphGibberella<br />

fujikuroi (Sawada) Ito in Ito & K. Kimura] and its association with maize.<br />

Thus, species of Fusarium endophytes include those fungi that occupy the<br />

intercellular spaces of plants, and the intercellular infections may be localized<br />

to roots. However, localization to roots is not mutually exclusive as<br />

some Fusarium species, while living as root endophytes, may also infect<br />

above ground plant organs, although the foliage origin of such hyphae<br />

from the endophytic infections in the roots has not been established for<br />

all species. Indeed, secondary infections from aerial spores are suspected<br />

of contributing to most foliage infection (Adams 1921; Boshoff et al. 1996;<br />

Kang and Buchenauer 2000) and, as discussed below, there is the possibility<br />

of different specific fungal strains that infect specific tissue types, especially<br />

the flower infections.<br />

8.2<br />

Plant and Fungus Interactions<br />

Contrary to the association found in Neotyphodium grass species, Fusarium<br />

species are not necessarily obligate endophytes. Indeed, as presented


8RootEndophyticFusarium Species 135<br />

Fig.8.1. a–e Light micrographs of the endophytic habit of a non-virulent isolate of Fusarium<br />

verticillioides, RRC 826, in roots of 2-week-old maize seedlings. a Hyphae (arrowhead)<br />

running parallel within intercellular spaces of the first internodes and junction of primary<br />

root of maize (54.4X). b Higher magnification of maize root showing a branching septate<br />

hypha (arrowhead) between two cell walls (272X, phase contrast). c Hyphae running parallel<br />

within intercellular spaces with a branching hypha at arrowhead (272X, phase contrast).<br />

d Cross section of a secondary root with hyphae (arrowhead) in the intercellular spaces<br />

(54.5X) (very dark spherical bodies within cells are artifacts of staining). e Cross section<br />

through the cortex of a primary root with groups of hyphae (arrowheads) in the intercellular<br />

spaces (109X, phase contrast) (from Bacon and Hinton 1996, with permission)<br />

below, their nutritional physiology, i.e., parasitic and saprophytic, predicts<br />

a transitory nature of the endophytic phase of Fusarium associations.<br />

Understanding to what extent species or isolates of this genus are endophytic<br />

is hampered by studies, histological or otherwise, that inadequately<br />

describe the qualitative and quantitative distribution of Fusarium within<br />

hosts. Nor are there studies to indicate any host requirements or benefits<br />

derived from such association that are indicated and characteristic of<br />

those derived from the Neotyphodium grassendophytes(forreview,see<br />

Bacon and White 2000). However, some studies are highly suggestive of<br />

benefits, at least to the fungus (Lee et al. 1995; Yates et al. 1997; Munkvold<br />

and Carlton 1997; Kuldau and Yates 2000; Pinto et al. 2000; Bacon et al.<br />

2004).


136 C.W. Bacon, I.E. Yates<br />

ReportsabouttheinteractionsofF. verticillioides with maize have been<br />

contradictory since the initial description of this fungus as both a pathogenic<br />

and a symptomless infection (Sheldon 1902; Voorhees 1934). Anatomical<br />

features of pathogenic infections by Fusarium species, including F. verticillioides,<br />

have been reviewed (Pennypacker 1981; Bacon and Hinton 1996),<br />

although some of the reports were concerned mainly with above ground<br />

plant parts. Voorhees (1934) described an initial infection of F. verticillioides<br />

intorootsthatoccurredfromthesoil.Infectiontookplaceviathe<br />

primary radicle by the fungus entering the epidermis, although it can also<br />

enter through ruptures produced in the cortex by emerging lateral roots.<br />

Since the endodermis of the young radicle acts as the barrier against penetration<br />

(Voorhees 1934), spread of infection into the stele is prevented.<br />

However, infection of the stele can occur from soil via the wounds produced<br />

by adventitious and lateral roots, suggesting that the degree and rate<br />

ofsuberizationinyoungseedlingscanserveasthekeytothenatureof<br />

disease development as opposed to development and the duration of the<br />

symptomless state.<br />

Modern day maize cultivars are more resistant and, thus, symptomless<br />

endophytic colonization by F. verticillioides is the rule today (Foley 1962;<br />

Kommedahl and Siggerirsson 1975; Thomas and Buddenhagen 1980; Bacon<br />

and Hinton 1988; Ayers et al. 1989; Leslie et al. 1990; Corell et al. 1992;<br />

Bentley et al. 1995; Ahmed et al. 1996; Anaya and Roncero 1996; Bacon and<br />

Hinton 1996; Bai 1996; Bakan et al. 2002; Anjaiah et al. 2003). Maize kernels<br />

are universally infected by F. verticillioides and related species, but disease<br />

symptoms are rarely exhibited. Possibly, plant breeding and selection may<br />

also form the basis for symptomless root infection in other agricultural<br />

species and cultivars. Further, species and strain genetics are important<br />

within the overall population of each species, and are important in the<br />

overall impact of a species on a host. Indeed, there is now information to<br />

indicate that a genetic change that will convert a Fusarium pathogen to<br />

a nonpathogenic endophytic mutualist can occur (Freeman and Rodriguez<br />

1993). The following are brief descriptions of both the symptomless and<br />

pathogenic infections.<br />

Molecular tools have also been utilized to study the association of F. verticillioides<br />

with maize and other plants. In situ studies of this species were<br />

made possible with avirulent isolates of F. verticillioides transformed with<br />

a plasmid containing the gusA gene coding for β-glucoronidase (GUS)<br />

and the hygr gene coding for hygromycin resistance (Yates et al. 1999).<br />

Subsequent GUS activity is detectable by histochemical and fluorometric<br />

enzymatic assays during the colonization period. The results indicated<br />

that F. verticillioides could be traced from the initial seed, to recovery<br />

from roots of plants produced from these seed, up through the stem, and<br />

finally isolated internally from seed (Bacon et al. 2001). Further, it was


8RootEndophyticFusarium Species 137<br />

Fig.8.2. a–e Scanning electron micrographs of a symptomless F. verticillioides-infected<br />

maize kernel. a A longitudinal section showing the location of the fungus at the tip cap<br />

(x13). b Magnified version of a, showing the location of the fungus in box (arrow) below<br />

the vascular tissues (v). c AhyphaofF. verticillioides (arrow) within the boxed area<br />

of b. d Growth of the fungus on culture medium showing chains of microconidia (arrow)<br />

characteristic of this species. e Growth of the fungus on the other half of the sectioned maize<br />

kernel in a, above, incubated aseptically for 2 days on damp filter paper, and showing the<br />

chains of microconidia, which established that the hypha in a was that of F. verticillioides<br />

(from Bacon et al. 1992, with permission)<br />

demonstrated that seed produced from these plants are sound, the fungus<br />

is internally seed borne (Fig. 8.2a–f), and produce seedlings that are infected<br />

with the transformed fungus (Bacon et al. 2001). Thus, this species is<br />

disseminated vertically, but horizontal dissemination is expected to occur<br />

through wounds due to the activity of insects (Munkvold and Carlton 1997),<br />

from soil to roots and other injured plant parts (Foley 1962), as well as via<br />

aerial borne spores. Most Fusarium species are also disseminated horizontally<br />

(Adams 1921), although vertical transmission is equally possible since


138 C.W. Bacon, I.E. Yates<br />

several species have been recovered from seed as endophytes (Gordon 1952;<br />

Fisher and Petrini 1992).<br />

8.2.1<br />

Hemibiotrophic Characteristics<br />

Hemibiotrophic fungi include those that infect living tissue, similar to<br />

biotrophs, but after an extended incubation period of days or weeks the<br />

infected tissue dies, within which the fungus now continues to develop as<br />

a saprotroph, usually resulting in sporulation (Luttrell 1974). The distinction<br />

between a hemibiotrophic parasite and a necrotrophic parasite is one<br />

of tissue infections. Necrotrophs kill host tissue in advance of penetration;<br />

while biotrophs penetrate, infect and obtain their food from living tissue,<br />

and this includes sporulation on living tissue. Having made the distinction<br />

between the basic terminologies used to distinguish the nutrition association<br />

of parasitic fungi with vascular plants, we now can apply this to<br />

endophytic Fusarium species.<br />

Following the definition above, Fusarium speciesshouldbeconsidered<br />

hemibiotrophs. That is, they are fungi that infect living tissue as biotrophs<br />

but after a latency period, which may last for a period of days to weeks, can<br />

cause host tissue to die, at which point the fungus becomes a saprotroph.<br />

Thus, endophytic Fusarium species, along with other fungal endophytes,<br />

are facultative biotrophic parasites, although this mode of nutrition can<br />

be a transient feature. Fungi belonging to Claviceps and Colletotrichum,as<br />

well as Ustilago maydis and Magnaporthe grisea,aretypicalhemibiotrophs.<br />

However, as we shall see below, not all isolates of the species F. verticillioides<br />

are hemibiotrophic and this might be due to either host and fungus genetics<br />

or environmental factors, or both. Certainly, in those situations where the<br />

saprotrophic stage is prevented, mycotoxin accumulation might be reduced,<br />

providing one point of reducing the concentration of these toxins.<br />

The parasitic association of Fusarium species with roots as endophytes<br />

may be viewed as a long-term event, perhaps confounded by plant breeding,<br />

and selection not for endophytic infection but for disease expression. Plant<br />

breeding might make it difficult to clearly define the degree of biotrophy<br />

characteristic for each Fusarium species relative to specific modern agricultural<br />

host cultivars that differ in disease resistance. In addition, the degree<br />

of biotrophy depends on the genetics of fungal strains and hosts, and the<br />

site of inoculation (Corell et al. 1992; Munkvold and Carlton 1997; Carter et<br />

al. 2000; Mesterhaszy et al. 2003). Thus, the association of Fusarium species<br />

with plants as symptomless endophytes cannot be explained entirely on<br />

the basis of the nutritional parasitic relationships described above. Complete<br />

understanding of Fusarium species as symptomless root endophytes


8RootEndophyticFusarium Species 139<br />

is hampered due to the complex of interactions, compounded by the fact<br />

that symptomless associations are altered by damaged portions of roots,<br />

due to predation by insects and wounds from emerging lateral roots. This<br />

results in a transformation of the symptomless state of the root-infecting<br />

species to the biotrophic and saprophytic state that infects dying and dead<br />

tissue. Most Fusarium rot diseases, i.e., root rots, are characterized during<br />

this phase. However, this appearance of biotrophic and saprophytic states is<br />

not necessarily obligatory and, indeed, a plant may have the symptomless<br />

state in roots while other parts of the same plant may be diseased. It is this<br />

aspect of the association that is of major concern in this review.<br />

8.2.2<br />

Histology<br />

Symptomless root infections are characteristic of many Fusarium species<br />

(Foley 1962; Pennypacker 1981; Wong et al. 1992; Leslie 1994; Bacon and<br />

Hinton 1996; Bowden and Leslie 1994; Gang et al. 1998; Nicholson et al.<br />

1998; Kedera et al. 1999; Kuldau and Yates 2000; Bai et al. 2002), and there<br />

are very aggressive or virulent strains of all species that serve as incitants<br />

of several plant diseases (Foley 1962; Malalasekera et al. 1973; Pennypacker<br />

1981; Manaka and Chelkowski 1985; Liddell and Burgess 1985; Wilcoxon et<br />

al. 1988; Fisher and Petrini 1992; Bacon and Hinton 1996; Bai 1996; Boshoff<br />

et al. 1996; Parry and Nicholson 1996; Kosiak et al. 1997; Nicholson et<br />

al. 1998; Wildermuth et al. 1999; Hysek et al. 2000; Ribichich et al. 2000).<br />

Disease development is considered to be a consequence of fungal and host<br />

genetics, while environmental biotic and abiotic factors negatively affect<br />

overall survival of the host (Schroeder and Christensen 1963; Bacon et al.<br />

1996; Ribichich et al. 2000).<br />

Detailed histological studies of specific Fusarium species are lacking<br />

since it is the disease state that is most often studied. Further, molecular<br />

analyses now suggest that these studies involved either different species or<br />

more than one species. For example, we now know that studies of scab or<br />

head blight of wheat consisted of several cryptic species. The flower-foliage<br />

disease or scab is caused primarily by F. graminearum and, to a lesser extent,<br />

by a new species F. pseudograminearum, while crown rot of wheat is caused<br />

by yet another new species, Gibberella coronicola (Aoki and O’Donnell<br />

1999; O’Donnell et al. 2000). Nevertheless, there are similarities and the<br />

following discussion takes these into account.<br />

Bacon and Hinton (1996) compared root and shoot infections by both<br />

virulent and non-virulent strains of F. verticillioides on maize using lightand<br />

electron-transmission-microscopy and concluded that this strain, and<br />

perhaps most others, should be considered as symptomless endophyte(s)


140 C.W. Bacon, I.E. Yates<br />

(Fig. 8.1a–e), especially since most isolates produce symptomless infections<br />

with most modern day cultivars of maize. It was also concluded that this<br />

species was not a vascular rot fungus, agreeing with the earlier observations<br />

of Pennypacker (1981) that F. verticillioides has the potential to be a cortical<br />

rot fungus (Fig. 8.1d–e). The hyphae of F. verticillioides run parallel within<br />

intercellular spaces, although branching hyphae are observed, especially<br />

in areas of branching roots (Fig. 8.1c). A study of the symptomless state<br />

in seedling roots suggests that the symptomless state persists beyond the<br />

seedling stage and contributes, without visual signs, to mycotoxin contamination<br />

of maize both before and during kernel development (Bacon and<br />

Hinton 1996). Symptomless root infections have been reported in plants<br />

infected by several species of Fusarium including F. graminearum (Gordon<br />

1952; Sieber et al. 1988), F. oxysporum f. sp. melonis (Katan 1971), F. oxysporum<br />

(Lemanceau et al. 1993), F. nivale, F. culmorum (Sieber et al. 1988),<br />

F. crookwellense (Boshoff et al. 1996), F. culmorum (Kang and Buchenauer<br />

1999); see additional species in the review of Kuldau and Yates (2000), and<br />

in Warren and Kommedahl (1973).<br />

Symptomless infection by Fusarium species varies and is related to the<br />

infection age of the hyphae (Baayen and Rykenbuerg 1999), and compartmentalization<br />

of the disease within resistant host tissues (Baayen et al. 1996)<br />

while others remain symptomless and biotrophic and are characterized as<br />

having interfacial membranes that separate fungal and plant plasma membranes<br />

(Fig. 8.3a–c) (Marchant 1966; Malalasekera et al. 1973; Baayen and<br />

Elgersma 1983; Bacon and Hinton 1996; Boshoff et al. 1996) The association<br />

is compatible over an extended time period, and non-specialized hyphae,<br />

which appear not to differ morphologically from the intercellular hyphae of<br />

symptomless infection (Bacon and Hinton 1996; Figs. 8.3a–c, Figs. 8.4a–f),<br />

are used for nutrient absorption. Endophytic hyphae of Fusarium spp. are<br />

not dormant (or quiescent), but are metabolically active throughout the<br />

association and within the intercellular spaces (Miller and Young 1985;<br />

Evans et al. 2000; Kang and Buchenauer 2000; Bacon et al. 2001). In such<br />

hyphae, there is a consistent lack of distinct nutrient absorbing structures<br />

characteristic of the usual pathogenic infections in rust and powdery<br />

mildews such as haustoria. Virulent strains of F. verticillioides do have intracellular<br />

hyphae that are not distinct haustoria (Fig. 8.4b–f) (Bacon and<br />

Hinton 1996). However, haustoria represent only one type of nutrient absorption<br />

structure. Fungal cell-wall-to-plant-wall appositions, as observed<br />

in most strains of F. verticillioides and other Fusarium speciesaswellas<br />

other fungal endophytes, are one of three fungus-plant cell nutrient absorbing<br />

structures (Honneger 1986). Thus, nutrient absorption by wall-to-wall<br />

contact and nutrients from the apoplasm are apparently just as efficient as<br />

intracellular hyphae since growth and colonization of the host is still accomplished,<br />

along with the production of secondary metabolites (Bacon et


8RootEndophyticFusarium Species 141<br />

Fig.8.3. a–c Transmission electron micrographs of the symptomless endophytic hyphae of<br />

F. verticillioides, RRC 826, as they appear in cross-section of 1- to 8-week-old plants of<br />

maize. a Two hyphae (f )inanintercellularspace(i)ofmaizecells(h); bar 1µm.b Another<br />

view of the intercellular nature of hyphae in roots of maize plants approximately 8 weeks<br />

old; bar 1µm.c Twogroupsofhyphaeinintercellularspacesoftherootcortex;bar 1µm<br />

(from Bacon and Hinton 1996, with permission)<br />

al. 2001). The accumulation of secondary compounds must require, a priori,<br />

a tremendous amount of energy. Certainly, the greater distribution<br />

of Fusarium hyphae in roots (Foley 1962; Warren and Kommedahl 1973;<br />

Kommedahl and Siggerirsson 1975; Blok and Bollen 1995; Bacon and Hin-


142 C.W. Bacon, I.E. Yates<br />

Fig.8.4. a–f Transmission electron micrographs of a virulent strain of Fusarium verticillioides,<br />

RRC pat, in maize plants during the early stages of seedling blight showing several<br />

variations of intracellular hyphae. a A noninfected plant showing intact chloroplast (ch);<br />

bar 1µm.b At 3 weeks, the fungus is primarily intracellular; chloroplasts, while intact, are<br />

becoming disorganized; one fungus (f ) is connected between cells with a hyphal penetration<br />

peg (arrowhead); bar 1µm.c Inter- and intra-cellular location of the fungus in leaves;<br />

intracellular fungus (f ) with an elongated hyphal penetration peg (arrowhead) within cell<br />

of maize (h); chloroplasts are no longer intact; bar 1µm.d Intercellular hyphae invading<br />

another cell in stem tissue of a 3-week-old plant; bar 1µm.e An intracellular hypha growing<br />

between two cells along and between the middle lamellae; bar 1µm.f Maize tissue showing<br />

extensive inter- and intra-cellular colonization; bar 1 µm (from Bacon and Hinton 1996,<br />

with permission)<br />

ton 1996) reflects the larger amounts of nutrients within the root apoplasm<br />

that accumulate as a sink from photosynthesis.<br />

Reports on symptomatic infections are numerous, although detailed histological<br />

studies of these types of infections tend to be restricted to the specific<br />

plant organs showing the effects (Kang and Buchenauer 2000; Boshoff<br />

et al. 1996; Ribichich et al. 2000). However, there are several histological<br />

similarities during the change from biotrophic phase to necrotrophic phase


8RootEndophyticFusarium Species 143<br />

(Adams 1921; Bennet 1931; Malalasekera et al. 1973; Wong et al. 1992; Bacon<br />

and Hinton 1996; Baayen and Rykenbuerg 1999; Kang and Buchenauer<br />

2000), reinforcing our concept that most Fusarium species are hemibiotrophic.<br />

Specialized intracellular infection and nutrient absorbing hyphae are<br />

absent during the biotrophic state of F. verticillioides infecting maize<br />

(Figs. 8.1a–e, 8.3a–c) and in other hosts as well (Malalasekera et al. 1973;<br />

Manaka and Chelkowski 1985; Honneger 1986; Boshoff et al. 1996; Kang<br />

and Buchenauer 1999, 2000), and in general these fall within the type 1<br />

category that ranges from the simple to the appressorium configuration<br />

(Honegger 1986). This type of fungus-plant cell interaction is characteristic<br />

of interactions that occur in different symbiotic systems (Honegger<br />

1986). We do not know if the lack of specialized nutrient absorption structures<br />

during this phase is characteristic of all Fusarium species. However,<br />

infection of host cells by specialized hyphae is described for F. culmorum<br />

and other species during the change to the intracellular stage of infection<br />

(Malalasekeraet al. 1973; Kang and Buchenauer 2000) and these would fall<br />

within the definition of one of the type 2 intracellular haustoria without<br />

sheath and papilla described by Honegger (1986) as indicative of mutualistic<br />

symbioses commonly found in lichens. However, the nature of the<br />

association with a particular host may vary, as some endophytic associations<br />

appear to be more epicuticular (on glumes) than endophytic and<br />

systemic (Kang and Buchenauer 1999). A lack of specialized intracellular<br />

absorbing structures assures less injury to the host, thus insuring compatibility,<br />

and presumably nutrients are obtained from the apoplasm of the<br />

intercellular spaces, although there may be a fungus-directed source and<br />

sink relationship.<br />

8.2.3<br />

Mycotoxins<br />

Fusarium species are a highly successful group of fungi that produce a variety<br />

of secondary metabolites, some of which might be important in both<br />

the long- and short-term strategies of the species. Most biochemical studies<br />

have concentrated on the production of, and factors leading to the accumulation<br />

of, mycotoxins and related compounds. However, there are sporadic<br />

and often observational reports on the positive and negative effects of<br />

symptomless infections by Fusarium species on hosts, and on competing<br />

organisms, which may be activities of secondary metabolites. Fusarium<br />

species produce a wide diversity of mycotoxins, resulting in a variety of<br />

effects on animals (Marasas et al. 1984, 1988; Voss et al. 1990; Norred et<br />

al. 1992; Riley et al. 1993). Mycotoxins have been speculated to play an


144 C.W. Bacon, I.E. Yates<br />

important role in the long-term survival strategy of the producing fungus.<br />

However, in some instances, some mycotoxins might play an even greater<br />

role in the day-to-day competitive fitness strategies of Fusarium species<br />

and these are presented briefly below.<br />

8.2.4<br />

Mycotoxins and Host Relationships<br />

Fumonisins are mycotoxins and are the subject of current concern due to<br />

their widespread occurrence in maize and maize products. F. verticillioides<br />

and other fungi of the Liseola section produce fumonisins in maize and<br />

other commodities. Fumonisins were shown to accumulate in colonized<br />

maize roots early during maize seedling development, and more fumonisins<br />

are isolated from roots than from shoots at this early stage of growth<br />

(Bacon et al. 2001). Fumonisin mycotoxins probably also occur in the roots<br />

of other plant species, and a rooting response has been shown for one cultivar<br />

of tomato (Bacon and Williamson 1992). A genetic and morphological<br />

study of conidiation mutants of F. verticillioides yielded interesting results<br />

concerning the role of fumonisins in plants (Glenn et al. 2004). Wild type<br />

isolates produced enteroblastic phialidic conidia, while mutants incapable<br />

of enteroblastic conidiogenesis produced undulating germ tube-like outgrowths.<br />

These mutants were not capable of infecting maize roots, and<br />

varied in their fumonisin production. Although they could not infect the<br />

plant, only those mutants that could produce the fumonisin were able to<br />

cause death of seedlings, but only in a small sampling of maize cultivars.<br />

This suggests that fumonisins play a role in the pathogenic processes of<br />

some Fusarium species but expression of toxicity on the host depends on<br />

the host genotype.<br />

Other Fusarium mycotoxins that might have a function in the physiology<br />

of the association include those produced by F. graminearum and<br />

related species. These include deoxynivalenol (DON or vomitoxin, a type<br />

B trichothecene), other related trichothecenes, and zearalenone. The mycotoxin<br />

DON is by far the most dominant of the trichothecenes, occurring<br />

on oats, rye, and maize, and occasionally on rice, sorghum, and triticale.<br />

In addition to the economic impact of this toxin in reducing animal performance,<br />

DON has adverse effects on plant performance.<br />

DON has been implicated as a virulence factor for some hosts (Bai<br />

et al. 2002; Desjardins and Hahn 1997; Harris et al. 1999), although the<br />

concentrations produced may or may not directly correlate with the degree<br />

of fungal virulence (Desjardins et al. 1996; Carter et al. 2000; Bai et al.<br />

2001; Mesterházy et al. 2003). DON was isolated from florets and grains of<br />

rye, wheat, and maize (Desjardins et al. 1996), but nothing is known about


8RootEndophyticFusarium Species 145<br />

its distribution and early toxin production within root tissue, or about<br />

how much of this toxin is produced during any endophytic stage of wheat<br />

or maize root colonization by F. graminearum, the main cause of blight<br />

disease of cereals. Nevertheless, this species also colonizes maize stalks and<br />

all tissues of wheat asymptomatically (Sieber et al. 1988; Dodd 1992; Fisher<br />

et al. 1992). McLean (1996) has reviewed additional information on the<br />

phytotoxicity of various Fusarium metabolites.<br />

8.2.5<br />

Physiological Interactions and Defense Metabolites<br />

Positive physiological interactions with maize were recorded for several<br />

strains of F. verticillioides These include increased rooting (Bacon and<br />

Williamson 1992), and earlier lignification of roots in seedling plants (Yates<br />

et al. 1997). Biocontrol uses of several Fusarium species not only indicate<br />

the utility of the genus but also suggest possible mutualistic interactions<br />

derived from the associations, including insecticidal (Abado-Becognee et<br />

al. 1998), nematocidal (Hallmann and Sikora 1994a, 1994b), and fungicidal<br />

(Damicone and Manning 1982; Lemanceau et al. 1993) activities.<br />

The endophytic association of F. verticillioides with maize has evolutionary<br />

and physiological implications since it has been established that<br />

all maize isolates of this fungus can detoxify the host’s native antimicrobial<br />

compounds, the benzoxazinoids (Glenn and Bacon 1998; Glenn et al.<br />

2001, 2002). The ability to detoxify these compounds, the concentrations<br />

of which may be high in roots (Xie et al. 1991), has been interpreted as<br />

one mechanism by which endophytic colonization is not prevented, since<br />

the benzoxazinoids are especially toxic to fungi (Xie et al. 1991; Schulz and<br />

Wieland 1999; Sicker et al. 2000; Glenn et al. 2001, 2002). Further, strains of<br />

F. verticillioides isolated from banana cannot detoxify the benzoxazinoids<br />

(Glenn et al. 2001, 2002). Several other species of Fusarium can detoxify<br />

the benzoxazinoids, suggesting the importance of this mechanism to this<br />

group, as well as to other maize pathogens (Schulz and Wieland 1999), for<br />

the endophytic association with grasses (Glenn 2000; Sicker et al. 2000).<br />

It is well known that several species of bacteria are endophytic in plants,<br />

also occupying intercellular spaces [for review, see Chanway 1998; see also<br />

Chaps. 2 (Hallmann and Berg), 3 (Kloepper and Ryu), and 6 (Anand et al.)].<br />

Endophytic bacteria are ecological homologues for most species of endophytic<br />

Fusarium species, and compete for nutrients within the apoplasm.<br />

Several biocontrol strategies are based on the use of bacterial endophytes<br />

[Chanway 1998; Kobayashi and Palumbo 2000; see Chaps. 3 (Kloepper and<br />

Ryu), and 4 (Berg and Hallmann)]. However, all Fusarium species examined<br />

produce fusaric acid (Bacon et al. 1996), and possibly other unknown an-


146 C.W. Bacon, I.E. Yates<br />

tibiotics that serve to control bacteria in planta. Fusaric acid is inhibitory<br />

to most strains of endophytic and rhizosphere bacteria (Notz et al. 2002;<br />

Bacon et al. 2004). Fusaric acid is moderately toxic to mammals (Porter<br />

et al. 1995), but is produced by most isolates of Fusarium (Bacon et al.<br />

1996). However, its activity is broad and it has pronounced effects on microorganisms.<br />

Fusaric acid is an antibiotic, showing activity against both<br />

Gram-negative and -positive bacteria, especially those used for biocontrol.<br />

Fusaric acid was shown to interact with genes specifically used by the<br />

biocontrol bacterium Pseudomonas fluorescens by preventing expression<br />

of genes encoding specific inhibitors of fungi (Notz et al 2002). A similar<br />

mode of action has been established for DON, which prevents the production<br />

of the chitinase gene expressed by the biocontrol agent Trichoderma<br />

atroviride (Lutz et al. 2003).<br />

Knowledge of the complete spectra of activity for most Fusarium secondary<br />

metabolites is incomplete. Moniliformin is a weak mycotoxin and<br />

an even better phytotoxin for specific hosts (Cole et al. 1973), but any overall<br />

effects pertaining to Fusarium species are unknown. The beauvericins<br />

show strong selective insecticidal properties (Vesonder and Hesseltine<br />

1981; Plattner and Nelson 1994; Logrieco et al. 1998). The beauvericins<br />

are produced by at least 12 Fusarium species (Logrieco et al. 1998), and<br />

can protect Fusarium-infected plants from insect predation. The chemical<br />

isolation and some biological activities of numerous other metabolites<br />

have been reviewed by Marasas et al. (1984) and Thrane (1989) and include<br />

the enniatins, butenolide, wortmannin, diacetoxyscripenol, nivalenol, visoltricin,<br />

chrysogine, culmorin, aurofusarin, equisetin, fusoproliferatum,<br />

fusarochromanone, acuminatopyrone, fusamarin, chlamydosporol, additional<br />

derivatives of T-2 toxins, and several unidentified antibiotics. The<br />

identities of the Fusarium speciesproducingthesemetabolitesandamore<br />

comprehensive listing of the activities of these and other mycotoxins can<br />

be obtained from Thrane (1989, 2001) and Summerell et al. (2001).<br />

8.3<br />

Summary<br />

Fusarium is a very important genus from the point of view of food production<br />

and food safety. Fusarium species exist as intercellular root endophytes<br />

in both cultivated and wild plants and their role during the symptomless<br />

state of infection is ambiguously defined. However, many species are<br />

pathogenic, causing diseases such as root, stem, and ear rot on crop plants,<br />

thereby reducing plant productivity. F. verticillioides and other Fusarium<br />

species are unique endophytes, with similarities to other endophyte species<br />

such as the foliage endophytes of forage grasses, but are more versatile since


8RootEndophyticFusarium Species 147<br />

they are also hemibiotrophic in their associations with plants. The problems<br />

associated with this genus are global as their distribution is worldwide,<br />

and most host plants are susceptible to infection by one or more<br />

Fusarium species. Certain biotic and abiotic factors may alter Fusarium<br />

relationships with plants from a symptomless endophytic association to<br />

a hemibiotrophic and finally a saprotrophic association where mycotoxins<br />

might accumulate, leading to animal and human health concerns. To summarize,<br />

a major concern is that many Fusarium species produce mycotoxins<br />

that are harmful to humans and animals ingesting food or feed products colonized<br />

by the fungus. The mycotoxins are produced during the pathogenic<br />

and saprophytic states, but the infections caused by these two states can be<br />

initiated from the symptomless root endophytic biotrophic state.<br />

The dual characterization of F. verticillioides as both a pathogen and<br />

a symptomless endophyte indicates both the complex relationship of this<br />

species with plants as well as suggesting similar complexities for other<br />

Fusarium-plant interactions. Consequently, the development of appropriate<br />

control measures for virulent Fusarium isolates are expected to be<br />

difficult. For example, on the one hand, diseases and mycotoxins produced<br />

by F. verticillioides must be controlled, while on the other hand, the intimate<br />

association of the endophytic state of this species appears to confer some<br />

positive competitive fitness traits to certain plants. The extent to which this<br />

occurs due to present-day plant breeding efforts will be determined with<br />

more detailed studies. The association of this genus with roots as symptomless<br />

endophytes indicates a role for these fungi in nutrition, and suggests<br />

the importance of the root as an endophytic niche during the co-evolution<br />

of Fusarium species and plants.<br />

<strong>References</strong><br />

Abado-Becognee K, Fleurat-Lessard F, Creppy EE, Melcion D (1998) Effects of fumonisin B1<br />

on growth and metabolism of larvae of the yellow mealworm, Tenebrio molitor.Entomol<br />

Exp Appl 86:135–143<br />

Aoki T, O’Donnell K (1999) Morphological and molecular characterization of Fusarium<br />

pseudograminearum sp nov. formerly recognized as the Group 1 population of F. graminearum.<br />

Mycologia 91:597–609<br />

Adams JE (1921) Observation on wheat scab in Pennsylvania and its pathological histology.<br />

Phytopathology 11:115–124<br />

Ahmed KZ, Mesterhazy A, Bartok T, Sagi F (1996) In vitro techniques for selecting wheat<br />

(Triticum aestivum L) for Fusarium-resistance 2. Culture filtrate technique and inheritance<br />

of Fusarium-resistance in the somaclones. Euphytica 91:341–349<br />

Anaya N, Roncero MIG (1996) Stress-induced rearrangement of Fusarium retrotransposon<br />

sequences. Mol Gen Genet 253:89–94<br />

Anjaiah V, Cornelis P, Koedam N (2003) Effect of genotype and root colonization in biological<br />

control of fusarium wilts in pigeonpea and chickpea by Pseudomonas aeruginosa PNA1.<br />

Can J Microbiol 49:85–91


148 C.W. Bacon, I.E. Yates<br />

Ayers JE, Nelson PE, Krause RA (1989) Fungi associated with maize stalk rot in Pennsylvania<br />

in 1070 and 1971. Plant Dis Rep 56:836–839<br />

Baayen RP, Elgersma DM (1983) Colonization and histopathology of susceptible and resistant<br />

carnation cultivars infected with Fusarium oxysporum f.sp. dianthi. NethJPlant<br />

Pathol 91:119–135<br />

Baayen RP, Rykenbuerg FHJ (1999) Fine structure of the early interaction of lily roots with<br />

Fusarium oxysporum f.sp. lilii. Eur J Plant Pathol 105:431–443<br />

Baayen RP, Ouellette GB, Rioux D (1996) Compartmentalization of decay in carnations<br />

resistant to Fusarium oxysporum f.sp. dianthi. Phytopathology 86:1018–<br />

1031<br />

Backhouse D, Burgess LW, Summerell BA (2001) Biogeography of Fusarium. In:SummerellBA,LeslieJF,BackhouseD,BrydenWL,BurgessLW(eds)Fusarium<br />

Paul E. Nelson<br />

Memorial Symposium APS Press, St. Paul, MN, pp 122–137<br />

Bacon CW, Hinton DM (1988) Pathogenic differences in ten isolates of Fusarium moniliforme<br />

Sheldon. Mycol Soc Am Newslett 39:S19–S20<br />

Bacon CW, Hinton DM (1996) Symptomless endophytic colonization of maize by Fusarium<br />

moniliforme. Can J Bot 74:1195–1202<br />

Bacon CW, White JF Jr (eds) (2000) Microbial endophytes. Dekker, New York<br />

Bacon CW, Williamson JW (1992) Interactions of Fusarium moniliforme, its metabolites<br />

and bacteria with maize. Mycopathologia 117:65–71<br />

Bacon CW, Lyons PC, Porter JK, Robbins JD (1986) Ergot toxicity from endophyte-infected<br />

grasses: a review Agron J 78:106–116<br />

Bacon CW, Porter JK, Norred WP, Leslie JF (1996) Production of fusaric acid by Fusarium<br />

species. Appl Environ Microbiol 62:4039–4043<br />

Bacon CW, Yates IE, Hinton DM, Meredith R (2001) Biological control of Fusarium moniliforme<br />

in maize. Environ Health Perspect 109:325–332<br />

Bacon CW, Hinton DM, Porter JK, Glenn AE, Kuldau GA (2004) Fusaric acid, a Fusarium<br />

verticillioides metabolite, antagonistic to the endophytic biocontrol bacterium Bacillus<br />

mojavensis. Can J Bot 82:878–885<br />

Bai GH (1996) Variation in Fusarium graminearum and cultivar resistance to wheat scab.<br />

Plant Dis 80:975–979<br />

Bai GH, Plattner R, Desjardins A, Kolb F (2001) Resistance to Fusarium head blight and<br />

deoxynivalenol accumulation in wheat. Plant Breeding 120:1–6<br />

Bai GH, Desjardins AE, Plattner RD (2002) Deoxynivalenol-nonproducing Fusarium<br />

graminearum causes initial infection, but does not cause disease spread in wheat spikes.<br />

Mycopathologia 153:91–98<br />

Bakan B, Melcion D, Richard-Molard D, Cahagnier B (2002) Fungal growth and Fusarium<br />

mycotoxin content in isogenic traditional maize and genetically modified maize grown<br />

in France and Spain. J Agric Food Chem 50:728–731<br />

Bentley S, Pegg KG, Dale JL (1995) Genetic variation among a world-wide collection of<br />

isolates of Fusarium oxysporum fspcubense analysed by RAPD-PCR fingerprinting.<br />

Mycol Res 99:1378–1384<br />

Bergman BHH, Bakker-Van der Voort MAM (1979) Latent infection of tulip bulbs by<br />

Fusarium oxysporum. Neth J Plant Pathol 85:187–195<br />

Bennett FT (1931) Gibberella saubinetti (Mont) Sacc. on British cereals. II. Physiological<br />

and pathological studies. Ann Appl Biol 18:158–177<br />

Blok WJ, Bollen GJ (1995) Fungi on the roots and stem bases of asparagus in the Netherlands:<br />

species and pathogenicity. Eur J Plant Pathol 101:15–24<br />

Boshoff WHP, Pretorius ZA, Swart WJ, Jacobs AS (1996) A comparison of scab development<br />

in wheat infected by Fusarium graminearum and Fusarium crookswellense.Phytopathology<br />

86:558–563


8RootEndophyticFusarium Species 149<br />

Bowden RL, Leslie JF (1994) Diversity of Gibberella zeae at small spatial scales. Phytopathology<br />

86:1140<br />

Carter JP, Rezanoor HN, Desjardins AE, Nicholson P (2000) Variation in Fusarium graminearum<br />

isolates from Nepal associated with their host of origin. Plant Pathol 49:452–460<br />

Chanway CP (1998) Bacterial endophytes: ecological and practical implications. Sydowia<br />

50:149–170<br />

Cole RJ, Kieksey JW, Cutler HG, Doupnik BL, Peckham JC (1973) Toxin from Fusarium<br />

moniliforme: effects on plants and animals. Science 179:1324–1326<br />

Corell JC, Gordon TR, McCain AH (1992) Genetic diversity in California and Florida populations<br />

of the pitch canker fungus Fusarium subglutinans. f.sp.pina Phytopathology<br />

82:415–420<br />

Damicone JP, Manning WJ (1982) Avirulent strains of Fusarium oxysporum protect asparagus<br />

seedling from crown rot. Can J Plant Pathol 4:143–146<br />

Desjardins AE, Hahn TM (1997) Mycotoxins in plant pathogenesis. Mol Plant Microbe<br />

Interact 10:147–152<br />

Desjardins AE, Proctor RH, Bai GH, McCormick SP, Shaner G, Buechley G, Hahn TM (1996)<br />

Reduced virulence of trichothecenes antibiotics-nonproducing mutants of Gibberella<br />

zeae in wheat field tests. Mol Plant Microbe Interact 9:775–781<br />

Diehl WW (1950) Balansia and the Balansiae in America. In: USDA Agric Monogr 4, US<br />

Govt Print Office, Washington, DC, pp 1–82<br />

Dodd JL (1992) The role of plant stresses in development of maize stalk rots. Plant Dis<br />

64:533–537<br />

Evans CK, Xie W, Dill-Macky R, Mirocha CJ (2000) Biosynthesis of deoxynivalenol in<br />

spikelets of barley inoculated with macroconidia of Fusarium graminearum. PlantDis<br />

84:654–660<br />

Fisher PJ, Petrini O (1992) Fungal saprobes and pathogens as endophytes of rice (Oryza<br />

sativa) New Phytol 120:137–143<br />

Fisher PJ, Petrini O, Scott HML (1992) The distribution of some fungal and bacterial<br />

endophytes in maize (Zea mays L.). New Phytol 122:299–305<br />

Foley DC (1962) Systemic infection of corn by Fusarium moniliforme. Phytopathology<br />

52:870–872<br />

Freeman S, Rodriguez RJ (1993) Genetic conversion of a fungal plant pathogen to a nonpathogenic,<br />

endophytic mutualist. Science 260:75–78<br />

Gang G, Miedaner T, Schuhmacher U, Schollenberger M, Geiger HH (1998) Deoxynivalenol<br />

and nivalenol production by Fusarium culmorum isolates differing in aggressiveness<br />

toward winter rye. Phytopathology 88:879–884<br />

Glenn AE, Bacon CW (1998) Detoxification of benzoxazinoids by Fusarium moniliforme<br />

and allies. Phytopathology 88:S–S32<br />

Glenn AE, Hinton DM, Yates IE, Bacon CW (2001) Detoxification of corn antimicrobial<br />

compounds as the basis for isolating Fusarium verticillioides and some other Fusarium<br />

species from corn. Appl Environ Microbiol 67:2973–2981<br />

Glenn AE, Gold SE, Bacon CW (2002) Fdb1 and Fdb2, Fusarium verticillioides loci necessary<br />

for detoxification of preformed antimicrobials from corn. Mol Plant Microbe Interact<br />

15:91–101<br />

Glenn AE, Richardson EA, Bacon CW (2004) Genetic and morphological characterization<br />

of a Fusarium verticillioides conidiation mutant. Mycologia 96:968–980<br />

Gordon WL (1952) The occurrence of Fusarium species in Canada. II. Prevalence and<br />

taxonomy of Fusarium species in cereal seed. Can J Bot 30:209–251<br />

Hallmann J, Sikora RA (1994a) Influence of Fusarium oxysporum, a mutualistic fungal<br />

endophyte on Meloidogyne incognita infection of tomato. Pflanzenkr Pflanzenschutz<br />

101:475–481


150 C.W. Bacon, I.E. Yates<br />

Hallmann J, Sikora RA (1994b) Occurrence of plant parasitic nematodes and endophytic<br />

fungi in tomato plants in Kenya and their role as mutualistic synergists for biological<br />

control of root-knot nematodes. Int J Pest Manage 40:321–325<br />

Harris LJ, Desjardins AE, Plattner RD, Nicholson P, Butler G, Young JC, Weston G, Proctor<br />

RH, Hohn TM (1999) Possible role of trichothecene mycotoxins in virulence of<br />

Fusarium graminearum on maize. Plant Dis 83:954–960<br />

Honegger R (1986) Ultrastructural studies in lichens. II. Mycobiont and photobiont cell<br />

wall surface layers and adhering crystalline lichen products in flour parmeliaceae. New<br />

Phytol 103:797–808<br />

Hysek J, Vanova M, Sychrova E, Koutecka-Brozova J, Radova Z, Hajslova J (2000) Fusariosis<br />

of barley – the spectrum of species and the levels of mycotoxins (trichothecenes). Mitt<br />

Biol Bundesanst Land-Forstwirtsch 377:29<br />

Kang Z, Buchenauer H (1999) Immunocytochemical localization of fusarium toxins in<br />

infected wheat spikes by Fusarium culmorum. Physiol Mol Plant Pathol 55:275–288<br />

Kang Z, Buchenauer H (2000) Cytology and ultrastructure of the infection of wheat spikes<br />

by Fusarium culmorum. Mycol Res 104:1083–1093<br />

Katan J (1971) Symptomless carriers of the tomato wilt pathogen. Phytopathology 61:1213–<br />

1217<br />

Kedera CJ, Plattner RD, Desjardins AE (1999) Incidence of Fusarium spp and levels of<br />

fumonisin B1 in maize in western Kenya. Appl Environ Microbiol 65:41–44<br />

Kobayashi DY, Palumbo JD (2000) Bacterial endophytes and their effects on plants and uses<br />

in agriculture. In: Bacon CW, White JF Jr (eds) Microbial endophytes. Dekker, New York,<br />

pp 199–233<br />

Kommedahl T, Siggerirsson EI (1975) Prevalence of Fusarium species in roots and soil of<br />

grassland in Iceland. Mycologia 67:38–44<br />

Kosiak B, Torp M, Thrane U (1997) The occurrence of Fusarium spp. in Norwegian grain –<br />

a survey. Cereal Res Commun 25:595–506<br />

Kuldau GA, Yates IE (2000) Evidence for Fusarium endophytes in cultivated and wild plants.<br />

In: Bacon CW, White JF Jr (eds) Microbial endophytes. Dekker, New York, pp 85–117<br />

Lee JC, Lobkovsky E, Pliam NB, Strobel G, Clardy J (1995) Subglutinols A and B: immunosuppressive<br />

compounds from the endophytic fungus Fusarium subglutinans. JOrgChem<br />

60:7076–7077<br />

Lemanceau P, Bakker PAHM, Dekogel WJ, Alabouvette C, Schippers, B (1993) Antagonistic<br />

effect of nonpathogenic Fusarium oxysporum Fo47 and psuedobactin 358 upon<br />

pathogen Fusarium oxysporum f. sp. dianthi Appl Environ Microbiol 59:74–82<br />

Leslie JF, Pearson CAS, Nelson PE, Toussoun TA (1990) Fusarium spp. from corn,<br />

sorghum, and soybean fields in the central and eastern United States. Phytopathology<br />

80:343–350<br />

Liddell CM, Burgess LW (1985) Survival of Fusarium moniliforme at controlled temperature<br />

and relative humidity. Trans Br Mycol Soc 84:121–130<br />

Logrieco A, Moretti A, Castella G, Kostecki M, Golinski P, Ritieni A, Chelkowski J (1998)<br />

Beauvericin production by Fusarium species. Appl Environ Microbiol 64:3084–3088<br />

Luttrell ES (1974) Parasitism of fungi on vascular plants. Mycologia 66:1–15<br />

Lutz MP, Feichtinger G, Defago G, Duffy B (2003) Mycotoxigenic Fusarium and deoxynivalenol<br />

production repress chitinase gene expression in the biocontrol agent Trichoderma<br />

atroviride Pl. Appl Environ Microbiol 69:3077–3084<br />

Malalasekera RAP, Sanderson FR, Colhoun J (1973) Fusarium diseases of cereals. IX. Penetration<br />

and invasion of wheat seedlings by Fusarium culmorum and F. navale.TransBr<br />

Mycol Soc 60:453–462<br />

Manaka M, Chelkowski J (1985) Phytotoxicity and pathogenicity of Fusarium nivale towards<br />

cereal seedlings. Phytopathol Z 122:113–117


8RootEndophyticFusarium Species 151<br />

Marasas WFO, Nelson PE, Toussoum TA (1984) Toxigenic Fusarium species. Identity and<br />

mycotoxicology. The Pennsylvania State University Press, University Park, PA<br />

Marasas WFO, Kellerman TS, Gelderblom WCA, Coetzer JAW, Thiel PG, Van der Lugt JJ<br />

(1988) Leukoencephalomalacia in a horse induced by fumonisin B1 isolated from Fusarium<br />

moniliforme. Onderstepoort J Vet Res 55:197–203<br />

Marchant R (1966) Fine structure and spore germination in Fusarium culmorum.AnnBot<br />

30:821–830<br />

Mclean M (1996) The phytotoxicity of Fusarium metabolites: an update since 1989. Mycopathologia<br />

133:163–179<br />

Mesterházy A, Bartók T, Lamper C (2003) Influence of wheat cultivar, species of Fusarium,<br />

and isolate aggressiveness on the efficacy of fungicides for control of Fusarium head<br />

blight. Plant Dis 87:1107–1115<br />

Miller JD, Young CJ (1985) Deoxynivalenol in an experimental Fusarium graminearum<br />

infection of wheat. Can J Plant Pathol 7:132–134<br />

Munkvold GP, Carlton WM (1997) Influence of inoculation method on systemic Fusarium<br />

moniliforme infection of maize plants grown from infected seeds. Plant Dis<br />

81:211–216<br />

Nicholson P, Simpson DR, Weston G, Rezanoor HN, Lees AK, Parry DW, Joyce D (1998)<br />

Detection and quantification of Fusarium culmorum and Fusarium graminearum in<br />

cereals using PCR assays. Physiol Mol Plant Pathol 53:17–37<br />

Norred WP, Wang E, Yoo H, Riley RT, Merrill AH Jr (1992) In vitro toxicology of fumonisins<br />

and the mechanistic implications. Mycopathologia 117:73–78<br />

Notz R, Maurhofer M, Dubach H, Haas D, Defago G (2002) Fusaric acid-producing stains of<br />

Fusarium oxysjporum alter 2.4-diacetylpholoroglucinol biosynthetic gene expression in<br />

Psedomonas fluorescens CHAO in vitro and in the rhizosphere of wheat. Appl Environ<br />

Microbiol 68:2229–2235<br />

O’Donnell K, Kistler HC, McLeod L (2000) Gene genealogies reveal global phylogeographic<br />

structure and reproductive isolation among lineages of Fusarium graminearum, the<br />

fungus causing wheat scab. Proc Natl Acad Sci USA 97:7905–7910<br />

Parry DW, Nicholson P (1996) Development of a PCR assay to detect Fusarium poae in<br />

wheat. Plant Pathol 45:383–391<br />

Pennypacker BW (1981) Anatomical changes involved in the pathogenesis of plants by<br />

Fusarium. In: Nelson PE, Toussoun TA, Cook RJ (eds) Fusarium: diseases, biology, and<br />

taxonomy, Pennsylvania State University Press, University Park, PA, pp 400–408<br />

Pinto LSRC, Azevedo MD, Pereiro M, Viera MLC, Labate CA (2000) Symptomless infection<br />

of banana and maize by endophytic fungi impairs photosynthetic efficiency. New Phytol<br />

147:609–615<br />

Plattner RD, Nelson PE(1994) Production of beauvericin by a strain of Fusarium proliferatum<br />

isolated from corn fodder for swine. Appl Environ Microbiol 60:3894–3896<br />

Porter JK, Bacon CW, Wray EM, Hagler WM Jr (1995) Fusaric acid in Fusarium moniliforme<br />

cultures, corn, and feeds toxic to livestock and the neurochemical effects in the brain<br />

and pineal gland of rats. Nat Toxins 3:91–100<br />

Ribichich KF, Lopez SE, Vegetti AC (2000) Histopathological spikelet changes produced<br />

by Fusarium graminearum in susceptible and resistant wheat cultivars. Plant Dis<br />

84:794–802<br />

RileyRT,AnN-H,ShowkerJL,YooH-S,NorredWP,ChamberlainWJ,WangE,MerrillAHJr,<br />

Motelin G, Beasley VR, Haschek WM (1993) Alteration of tissue and serum sphinganine<br />

to sphingosine ratio: an early biomarker of exposure to fumonisin-containing feeds in<br />

pigs. Toxicol Appl Pharmacol 118:105–112<br />

Schroeder HW, Christensen JJ (1963) Factors affecting resistance of wheat to scab caused<br />

by Gibberella zeae. Phytopathology 53:831–838


152 C.W. Bacon, I.E. Yates<br />

Schulz M, Wieland I (1999) Variation in metabolism of BOA among species in various field<br />

communities – biochemical evidence for co-evolutionary processes in plant communities?<br />

Chemoecology 9:133–141<br />

Sheldon JL (1902) A corn mold (Fusarium moniliforme. sp.n.).AgricExpStnNebraska<br />

17:23–32<br />

Sicker D, Frey M, Schulz M, Gierl A (2000) Role of natural benzoxazinones in the survival<br />

strategy of plants. Int Rev Cytol 198:319–346<br />

Sieber T, Risen TK, Muller E, Fried PM (1988) Endophytic fungi in four winter wheat<br />

cultivars (Triticum aestivum L.) differing in resistance against Stagonospora nodorum<br />

(Berk.) Cast. Germ = Septorianodorum (Berk.) Berk. J Phytopathol 122:289–306<br />

Stone JK, Bacon CW, White JF Jr (2000) An overview of endophytic microbes: endophytism<br />

defined. In: Bacon CW, White JF Jr (eds) Microbial endophytes. Dekker, New York,<br />

pp 3–29<br />

Summerell BA, Leslie JF, Blackhouse D, Bryden WL, Burgess LW (2001) Fusarium Paul E.<br />

Nelson Memorial Symposium APS Press, St. Paul, MN<br />

Thomas MD, Buddenhagen IW (1980) Incidence and persistence of Fusarium moniliforme<br />

in symptomless maize kernels and seedlings in Nigeria. Mycologia 72:882–887<br />

Thrane U (1989) Fusarium species and their specific profiles of secondary metabolites.<br />

In: Chelkowski J (ed) Fusarium: mycotoxins, taxonomy and pathogenicity. Elsevier,<br />

Amsterdam, pp 199–225<br />

Thrane U (2001) Developments in the taxonomy of Fusarium species based on secondary<br />

metabolites. In: Summerell BA, Leslie JF, Blackhouse D, Bryden WL, Burgess LW (eds)<br />

Fusarium Paul E. Nelson Memorial Symposium APS Press, St. Paul, MN, pp 29–49<br />

Vesonder RF, Hesseltine CW (1981) Metabolites of Fusarium. In: Nelson PE, Toussoun TA,<br />

Cook RJ (eds) Fusarium: diseases, biology and taxonomy The Pennsylvania State University<br />

Press, University Park, PA, pp 350–364<br />

Voorhees RK (1934) Histological studies of a seedling disease of corn caused by Gibberella<br />

moniliformis. J Agric Res 49:1009–1015<br />

Voss KA, Plattner RD, Bacon CW, Norred WP (1990) Comparative studies of hepatotoxicity<br />

and fumonisin B1 and B2 content of water and chloroform/methanol extracts of<br />

Fusarium moniliforme strain MRC 826 culture material. Mycopathologia 112:81–92<br />

Warren HL, Kommedahl T (1973) Prevalence and pathogenicity to corn of Fusarium species<br />

from corn roots, rhizosphere, residues, and soil. Phytopathology 63:1288–1290<br />

Wilcoxon RD, Kommedahl T, Ozmon EA, Windels CE (1988) Fusarium species in scabby<br />

wheat from Minnesota and their pathogenicity to wheat. Phytopathology 78:586–589<br />

Wildermuth GB, McNamara RB, Sparks T (1999) Different expressions of resistance to<br />

crown rot in wheat. In: Magarey RC (ed) Proceedings of the First Australasian Soilborne<br />

Disease Symposium. Bureau of Sugar Experiment Stations, Brisbane, Australia, p 79<br />

Wong LSL, Tekauz A, Leslie D, Abramson D, McKenzie RIH (1992) Prevalence, distribution<br />

and importance of Fusarium head blight in wheat in Manitoba. Can J Plant Pathol<br />

14:233–238<br />

Xie YS, Arnason JT, Philogène BJR, Atkinson J, Morand P (1991) Distribution and variation<br />

of hydroxamic acids and related compounds in maize (Zea mays)rootsystem.CanJBot<br />

69:677–681<br />

Yates IE, Jaworski AJ (2000) Differential growth of Fusarium moniliforme relative to tissues<br />

from ‘Silver Queen’, a sweet maize. Can J Bot 78:472–480<br />

Yates IE, Bacon CW, Hinton DM (1997) Effects of endophytic infection by Fusarium moniliforme<br />

on corn growth and cellular morphology. Plant Dis 81:723–728<br />

Yates IE, Hiett KL, Kapczynski DR, Smart W, Glenn AE, Hinton DM, Bacon CW, Meinersmann<br />

R, Liu S, Jaworski AJ (1999) GUS transformation of the maize fungal endophyte<br />

Fusarium moniliforme. Mycol Res 103:129–136


9<br />

Microbial Endophytes of Orchid Roots<br />

Paul Bayman, J. Tupac Otero<br />

9.1<br />

Introduction<br />

Orchids are interesting on many levels. Their beautiful and bizarre adaptations<br />

for pollination have fascinated many people, including Darwin (1887).<br />

Interest in orchid pollination biology has overshadowed another, equally<br />

important symbiosis: mycorrhizal relationships. In turn, interest in orchid<br />

mycorrhizae has overshadowed the relationships between orchids and endophytes.<br />

In most cases, studies on orchid root fungi have ignored all fungi<br />

not thought to be mycorrhizal. In some cases, fungi in orchid roots have<br />

been presumed to be mycorrhizal when they may in fact be endophytes.<br />

Thus the frequency, diversity and importance of orchid root endophytes<br />

remain largely unexplored.<br />

The goal of this chapter is to review the incidence and importance of<br />

endophytes in orchid roots, and to disentangle the literature on mycorrhizal<br />

fungi from that on non-mycorrhizal endophytes.<br />

9.2<br />

Habits and Types of Orchid Roots<br />

The Orchidaceae is one of the largest families of plants, with 25,000 species –<br />

close to one-tenth of all known flowering plant species (Dressler 1990).<br />

Orchids are found in all except the most extreme terrestrial environments.<br />

Orchids are epiphytic (growing on plants), lithophytic (growing on<br />

rocks), terrestrial, or, in a few cases, some combination thereof. Epiphytic<br />

and lithophytic orchids are all tropical or subtropical; they comprise about<br />

75% of all species (Dressler 1990). Terrestrial orchids are found worldwide;<br />

all temperate orchids are terrestrial. Most of what is known about<br />

orchid-fungal associations comes from terrestrial orchids (perhaps because<br />

Paul Bayman: Departamento de Biologia, Universidad de Puerto Rico – Rio Piedras, PO Box<br />

23360, San Juan, PR 00931, USA, E-mail: pbayman@upracd.upr.clu.edu<br />

J. Tupac Otero: Universidad Nacional de Colombia-Palmira, Departamento de Ciencias<br />

Agrícolas, AA 237, Palmira, Valle del Cauca, Colombia<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


154 P. Bayman, J.T. Otero<br />

most orchidologists live in temperate countries) even though they are less<br />

speciose than epiphytes.<br />

Roots of epiphytic/lithophytic and terrestrial orchids differ (Rasmussen<br />

1995). Epiphytic and lithophytic roots are ecologically equivalent because<br />

inbothcasestherootsareexposedtolightandair.Rootsofepiphytic<br />

and lithophytic orchids are photosynthetic, perennial, and fairly constant<br />

throughout the year. Roots of terrestrial orchids, in contrast, are usually<br />

non-photosynthetic, live ≤ 3 years, and often show marked seasonal differences<br />

in growth and composition. They are usually buried in soil or<br />

leaf litter. Some terrestrial orchids have two morphologically distinct types<br />

of roots, one of which is mycorrhizal (Rasmussen 1995). Non-mycorrhizal<br />

roots tend to have more xylem and more amyloplasts than mycorrhizal<br />

roots.<br />

Orchid roots have a velamen, a multiple epidermis of one to several layers<br />

of thin-walled cells. The velamen helps the root trap water and probably<br />

nutrients (Dressler 1990; Rasmusssen 1995). It has been suggested that the<br />

velamen evolved to facilitate colonization of roots by mycorrhizal fungi<br />

(Dressler 1990); this seems unlikely because pelotons are more common in<br />

the cortex than in the velamen. (Pelotons are coils of hyphae within root<br />

cells, and are characteristic of orchid mycorrhizae). Also, epiphytic orchid<br />

roots generally have more developed velamens than terrestrial orchid roots<br />

(Dressler 1990), even though the frequency of mycorrhizal infection is<br />

lower.<br />

Terrestrial orchids are usually obligately mycorrhizal, even as adults<br />

(Rasmussen 1995). Some species are non-photosynthetic (or more accurately,<br />

myco-heterotrophic) and depend on fungi for nutrition (see<br />

Sect. 9.8). Many epiphytic orchids are facultatively mycorrhizal, at least<br />

as adults, and there is wide variation in frequency of mycorrhizal colonization;<br />

the same is probably true of epilithic orchids (Hadley and Williamson<br />

1972; Lesica and Antibus 1990; Goh et al. 1992; Richardson et al. 1993;<br />

Zelmer et al. 1996; Currah et al. 1997; Bayman et al. 1997; Rivas et al. 1998;<br />

Otero et al. 2002; Rasmussen 2002).<br />

9.3<br />

Bacteria as Epiphytes and Endophytes of Orchid Roots<br />

In general, bacterial root endophytes have been studied in an agricultural<br />

context, rather than in an ecological or biodiversity context [Chanway 1995;<br />

see Chaps. 6 (Anand et al.) and 19 (van Overbeek et al.)]. Orchids are no<br />

exception: interest in bacteria in orchid roots has focused on pathogens<br />

of cultivated plants rather than on endophytes of wild plants (Hadley et<br />

al. 1987). This oversight is unfortunate, since endophytic bacteria may be


9 Microbial Endophytes of Orchid Roots 155<br />

importantforthehealthofwildplantsaswellascropplants(Hallmannet<br />

al. 1997; Sturz et al. 2000).<br />

Most knowledge of endophytic bacteria in orchid roots comes from<br />

studies in Australia. Endophytic bacteria were isolated from 12 species of<br />

terrestrial orchids in Western Australia (Wilkinson et al. 1994). The most<br />

common genus isolated was Pseudomonas,whichvariedfrom23%to73%<br />

of isolates from each orchid species. Most isolates from Pterostylis spp. were<br />

also Gram-negative, while most isolates from all other orchids were Grampositive.<br />

This difference may reflect morphological differences among the<br />

orchids: in Pterostylis the main absorptive organs are underground stems,<br />

whereas the other orchids tested use adventitious roots. Some of these<br />

bacteria stimulated germination in vitro of seeds of P. vittata (Wilkinson<br />

et al. 1989).<br />

Epiphytic bacteria on orchid roots have also been studied, mainly to<br />

determine if nitrogen-fixing bacteria can help orchids obtain nitrogen.<br />

Arthrobacter, Bacillus, Mycobacterium, Pseudomonas, Oscillatoria and Nostoc<br />

were isolated from the surfaces of roots of the terrestrial orchid Calanthe<br />

vestita,andBacillus, Curtobacterium, Flavobacterium, Nocardia, Pseudomonas,<br />

Rhodococcus, Xanthomonas and Nostoc were isolated from the<br />

surface of roots of the epiphyte Dendrobium (Tsavkelova et al. 2001). Of<br />

these, the cyanobacteria Oscillatoria and Nostoc are capable of nitrogen fixation;<br />

they were not isolated from soil collected near the plants, suggesting<br />

some special affinity for the root surface. These and other cyanobacteria<br />

formed a biofilm on the surface of roots of epiphytic orchids in a greenhouse<br />

(Tsavkelova et al. 2003). Cyanobacteria have also often been observed<br />

within velamen cells of epiphytic orchids (Dressler 1990; Sinclair 1990).<br />

Nitrogen-fixing bacteria also occur on epiphytic Tillandsia plants, which<br />

often occur together with orchids (Brighigna et al. 1992). Despite the interesting<br />

implications of these studies, the transfer of nitrogen from bacteria<br />

to orchid roots has not been demonstrated (Dressler 1990; Sinclair 1990).<br />

9.4<br />

Orchid Endophytes or Orchid Mycorrhizal Fungi?<br />

The focus of this chapter and volume is on endophytes, not mycorrhizal<br />

fungi. However, it is impossible to review the literature on orchid root<br />

endophytes without also discussing mycorrhizae. The study of orchid mycorrhizae<br />

and endophytes are so inextricably linked that orchids are a good<br />

example of how hard it can be to disentangle the two [see also Chaps. 14<br />

(Cairney) and 16 (Brundrett)].<br />

Mycorrhizae are mutualisms between plant roots and fungi, whereas endophytes<br />

are microorganisms growing inside plant tissues without causing


156 P. Bayman, J.T. Otero<br />

symptoms of disease (see Chap. 1 by Schulz and Boyle). This concept of<br />

mycorrhizae is functional and describes a relationship, whereas the concept<br />

of endophytes principally describes where an organism lives, without<br />

assuming or excluding the possibility of benefit for either party. This distinction<br />

is complicated by the fact that mutualisms are part of a continuum<br />

of symbiotic relationships, and a relationship that is mutualistic may become<br />

commensalistic or parasitic, or vice versa [Bronstein et al. 2003; see<br />

Chaps. 15 (Schulz) and 16 (Brundrett)]. It is further complicated by the<br />

fact that, unlike most mycorrhizae, orchid mycorrhizae are not known to<br />

provide any benefit to the fungal partner (Andersen and Rasmussen 1996,<br />

Taylor et al. 2002); in some cases and perhaps in all, the relationship is<br />

actually parasitic rather than mutualistic. Furthermore, fungal endophytes<br />

that are not mycorrhizal in the field may stimulate orchid seed growth<br />

in culture – ‘functional specificity’ as opposed to ‘ecological specificity’,<br />

as defined by Masuhara and Katsuya (1994). This means that seed germination<br />

and seedling growth tests in vitro may not be entirely accurate in<br />

distinguishing mycorrhizal fungi from non-mycorrhizal endophytes (Rasmussen<br />

2002). The most reliable criterion for mycorrhizae is the visual<br />

detection of pelotons in the root, but in tropical, epiphytic orchids, the<br />

pelotons observed are often degraded (Otero et al. 2002).<br />

The distinction between orchid mycorrhizal fungi and endophytes is<br />

further complicated by the inconsistent use of terminology in the literature.<br />

Most studies of Rhizoctonia-like fungi assume that the relationship<br />

is mycorrhizal without demonstrating any functional benefit to the plant.<br />

Since Rhizoctonia-like fungi can also be plant pathogens, endophytes or<br />

saprotrophs, in some cases this may be an unwarranted assumption (Alconero<br />

1969; Masuhara and Katsuya 1994; Carling et al. 1999; Rasmussen<br />

2002). A more precise (but less convenient) description would be ‘presumably<br />

mycorrhizal endophytes.’ On the other hand, some authors are aware<br />

of this problem and err on the side of caution, preferring to call their fungi<br />

‘endophytes’ for lack of functional evidence of a mycorrhizal relationship,<br />

even though they are almost certainly mycorrhizal (e.g., Hadley and Ong<br />

1978; Ramsay et al. 1986; Currah 1991; Currah et al. 1997; Richardson et<br />

al. 1993; Otero et al. 2002). So it is hard to say how many papers with<br />

‘mycorrhiza’ in the title really mean to say ‘endophyte’ and vice versa.<br />

However, when the papers that focus on mycorrhizal fungi are excluded,<br />

theexistingbodyofworkonorchidendophytesissurprisinglysmall<br />

(Currah et al. 1997). For example, an excellent, comprehensive treatise<br />

on orchids and their mycorrhizal relationships mentions non-mycorrhizal<br />

endophytes only in passing (Rasmussen 1995).<br />

To simplify things, in this chapter we will assume that Rhizoctonia-like<br />

fungi and their telomorphs (= sexual stages), including Ceratobasidium,<br />

Thanatephorus, Tulasnella, andSebacina, aremycorrhizalinassociations


9 Microbial Endophytes of Orchid Roots 157<br />

with orchids, that other basidiomycete fungi may be mycorrhizal in mycoheterotrophic<br />

orchids, and that members of other groups of fungi are<br />

endophytes and not mycorrhizal. Known exceptions to this generalization<br />

are mentioned where relevant.<br />

9.5<br />

Problems with the Taxonomy of Orchid Endophytic Fungi<br />

Three problems complicate the study of endophytic fungi, including orchid<br />

root endophytes. First, many endophytic fungi do not sporulate in pure culture,<br />

and efforts to induce sporulation in vitro are often unsuccessful. Since<br />

traditionally fungi are classified by their spores and spore-bearing structures,<br />

nonsporulating fungi are very difficult to identify. For this reason,<br />

unidentifiable fungi are often grouped into ‘morphospecies’ on the basis<br />

of colony color, morphology and growth rate on agar media (Gamboa and<br />

Bayman 2001). DNA sequencing studies have shown that this technique is<br />

quite successful at grouping related fungi together (Arnold et al. 2000; Lacap<br />

et al. 2003) – but they remain unnamed, making comparisons between<br />

studies very difficult.<br />

Second, many endophytic fungi are undescribed, and do not fit well into<br />

previously described taxa (Hawksworth and Rossman 1997; Hawksworth<br />

1991, 2000). This complicates the job of studying them, but it also provides<br />

an incentive to do so: endophytes may be a largely unstudied reservoir of<br />

fungal biodiversity (Hawksworth and Rossman 1997; Arnold et al. 2001).<br />

Third, some endophytes do not grow in culture. Culturing of microorganisms<br />

from plant tissues provides a skewed picture of the organisms that<br />

grow there (see Chap. 17 by Hallmann et al.). One solution to this problem<br />

is to use PCR-based methods to amplify DNA directly from orchid roots<br />

using fungal-specific primers. So far, such techniques have been used to<br />

study orchid mycorrhizal fungi (see Sect. 9.8), but not non-mycorrhizal endophytes.<br />

This approach has revealed that some endophytes of grass roots<br />

belong to previously unknown major taxa of fungi (Vandenkoornhuyse<br />

et al. 2002); it is possible that orchids also harbor important undescribed<br />

lineages of fungi However, these PCR-based approaches are also biased:<br />

specific PCR primers are needed to amplify the fungal DNA but not the<br />

plant DNA, and primers can preferentially amplify certain groups of fungi<br />

(Bruns et al. 1998); using more than one set of primers and varying amplification<br />

conditions may help reveal additional taxa.


158 P. Bayman, J.T. Otero<br />

9.6<br />

Host Specificity of Orchid Endophytes<br />

Host specificity of endophytes is important for estimates of global fungal<br />

biodiversity. There are many fungal species associated with each plant<br />

species, and if these fungi are host-specific, the number of endophyte<br />

species will increase linearly with the number of plant species. Ratios of fungal<br />

species to plant species are used to extrapolate fungal biodiversity from<br />

plant species richness; the most commonly cited ratio is 6:1 fungi: plants,<br />

including pathogens and endophytes (Hawksworth 2000). Endophytes are<br />

an undersampled group in terms of fungal biodiversity (Hawksworth 1991;<br />

Hawksworth and Rossman 1997; Fröhlich and Hyde 1999).There are three<br />

reasons to expect that studying orchid root endophytes may make a significant<br />

contribution to this biodiversity: (1) orchids represent almost 10% of<br />

angiosperm species; (2) orchid roots are anatomically, morphologically and<br />

ecologically different from other roots (see Sects. 9.2, 9.3) most orchids are<br />

in the tropics, probably the most undersampled area for fungal biodiversity<br />

(Fröhlich and Hyde 1999; Hawksworth 2000; Arnold et al. 2001).<br />

There is century-old debate about the specificity of orchids for mycorrhizal<br />

fungi, (see Arditti et al.1990; Rasmussen 1995, 2002; Taylor et al. 2002<br />

for reviews), which applies to non-mycorrhizal endophytes as well. However,<br />

few attempts have been made to compare non-mycorrhizal endophytes<br />

among orchid species using quantitative methods. Given the diversity of<br />

endophyticfungiinorchidrootsandthevariationinmethodsamongstudies,<br />

it is difficult to determine the levels of specificity and preference in the<br />

interaction. Taxonomic problems (see Sect. 9.5) also complicate the issue of<br />

host specificity in orchid mycorrhizal fungi: since many endophytes cannot<br />

be identified to the species level with confidence, it is difficult to determine<br />

whether different orchid species have different communities of endophytes.<br />

9.7<br />

Endophytic Fungi in Roots of Terrestrial,<br />

Photosynthetic Orchids<br />

Non-mycorrhizal endophytes have been isolated from various terrestrial,<br />

photosynthetic orchids (Table 9.1), most extensively by Randall Currah and<br />

associates in Canada (Table 9.1; see Currah et al. 1997; Taylor et al. 2002).<br />

The most common and widespread endophyte isolated is Phialocephala,<br />

one of the ‘dark septate endophytes.’ These fungi are also common in many<br />

plants [Currah et al. 1987; Fernando and Currah 1995; see Chaps. 7 (Sieber<br />

and Grünig), 12 (Girlanda et al.) and 15 (Schulz)]. They may function as


9 Microbial Endophytes of Orchid Roots 159<br />

Table 9.1. Non-Rhizoctonia fungireportedfromorchidrootsa . Basidiomycetes are in bold; those from myco-heterotrophic orchids have been<br />

assumed or shown to be mycorrhizal<br />

Orchids Fungi Location Reference<br />

Terrestrial,<br />

photosynthetic orchids<br />

Amerorchis rotundifolia Phialocephalab Canada Zelmer 1994<br />

Amerorchis rotundifolia Phialocephala fortinii Canada Currah et al. 1987<br />

Blettia striata Favolaschia thwaitesii Zambia Jonsson and Nylund 1979<br />

Calypso bulbosa Leptodontidium orchidicola, Phialocephala fortinii Canada Currah et al. 1987<br />

Coeloglossum viride Dactylella sp., Phialocephala Canada Zelmer 1994<br />

Coeloglossum viride Leptodontidium orchidicola, Trichosporiella multisporum Canada Currah et al. 1987<br />

Cymbidium sinense Mycena orchidicola sp.nov. China Fan et al. 1996<br />

Cypripedium calceolus Alternaria sp., Chaetomiumsp.Cylindrocarpon sp., Epicoccum purpureum, Canada Zelmer 1994<br />

Phialocephala, Phoma sp.<br />

Cypripedium candidum Acremonium killense, Humicola sp., Phialocephala Canada Zelmer 1994<br />

Cypripedium montanum Phialocephala Canada Zelmer 1994<br />

Cypripedium passerinum Phialocephala Canada Zelmer 1994<br />

Cypripedium reginae Fusarium sp. Canada Vujanovic et al. 2000<br />

Dactylorhiza majalis cf. Laccaria Denmark Kristiansen et al. 2001<br />

Epipactis microphyllac Tuber, other Pezizales France Selosse et al. 2004<br />

Epipactus helleborine Cylindrocarpon destructans, Humicola fuscoatra, Morchella sp.,<br />

Finland Salmia 1988<br />

Sordaria fimicola<br />

Goodyera oblongifolia Humicola sp., Phialocephala, Phoma sp., Thermomyces verrucosus Canada Zelmer 1994<br />

Listera cordata Penicillium sp., Phialocephala Canada Zelmer 1994<br />

Piperia unalascensis Phialocephala Canada Zelmer 1994<br />

Piperia unalascensis Sistotrema sp. Canada Currah et al. 1990


160 P. Bayman, J.T. Otero<br />

Table 9.1. (continued)<br />

Orchids Fungi Location Reference<br />

Platanthera dilatata Phialocephala Canada Zelmer 1994<br />

Platanthera hyperborea Acremonium kiliense, Phialocephala, Sporormia minima,<br />

Canada Zelmer 1994<br />

Thielavia basicola<br />

Platanthera hyperborea Leptodontidium orchidicola, Trichocladium opacum Canada Currah et al. 1987<br />

Platanthera obtusata Acremonium killense, Phialocephala Canada Zelmer 1994<br />

Platanthera obtusata Sistotrema sp. Canada Currah et al. 1990<br />

Platanthera praeclara Fusarium oxysporum, Phialocephala Canada Zelmer 1994<br />

Spiranthes lacera Acremonium kiliense, Phialocephala Canada Zelmer 1994<br />

Spiranthes magnicamporum Cylindrocarpon sp., Papulaspora sp., Phialophora richardsiae, Canada Zelmer 1994<br />

Ulocladium sp.<br />

Spiranthes romanzoffiana Acremonium kiliense, Cylindrocarpon sp., Gliomastix murorum, Canada Zelmer 1994<br />

Phialocephala<br />

Myco-heterotrophic orchids<br />

Cephalanthera austinae Thelephora-Tomentella (14 spp.) United States Taylor and Bruns 1997<br />

Corallorhiza maculata Armillaria melea United States Campbell 1970a<br />

Corallorhiza maculata Russulaceae (20 spp.) United States Taylor and Bruns 1997<br />

Corallorhiza maculata Russulaceae (3 spp.) United States Taylor and Bruns 1999<br />

Corallorhiza maculata Cylindrocarpon sp., Phialocephala Canada Zelmer 1994<br />

Corallorhiza maculata Leptodontidium orchidicola Canada Currah et al. 1987<br />

Corallorhiza striata Cylindrocarpon sp., Phialocephala Canada Zelmer 1994<br />

Corallorhiza trifida Phialocephala Canada Zelmer 1994<br />

Corallorhiza mertensiana Russulaceae (22 spp.) United States Taylor et al. 2003<br />

Corallorhiza striata Thelephora-Tomentella United States Taylor 1997<br />

Corallorhiza trifida Mycena thuja New Zealand Campbell 1970a


9 Microbial Endophytes of Orchid Roots 161<br />

Table 9.1. (continued)<br />

Orchids Fungi Location Reference<br />

Corallorhiza trifida yellow basidiomycete w/ clamps Canada Zelmer and Currah 1995<br />

Corallorhiza trifida Thelephora-Tomentella Scotland McKendrick et al. 2000<br />

Danhatchia australis<br />

Lycoperdon perlatum New Zealand Campbell 1970b<br />

(= Yoania australis)<br />

Didymoplexis minor Marasimius coniatus Paleotropics Burgeff 1959<br />

Galeola altissima Erythromyces crocicreas, Ganoderma australe,<br />

Hamada and Nakamura<br />

Loweporus tephroporus, Microporus affinus<br />

1963, Umata 1995<br />

Galeola (= Erythrorchis) ochobiensis Hymenochaete crocicreas Umata 1998<br />

Galeola (= Erythrorchis) ochobiensis Auricularia polytricha, Lyophyllum shimeji Umata 1997a, 1997b<br />

Galeola (= Erythrorchis) ochobiensis Lentinula edodes Umata 1998<br />

Galeola (= Erythrorchis) ochobiensis Lenzites betulinus, Trametes hirsuta Japan Umata 1999<br />

Galeola septentrionalis Armillaria mellea Japan Hamada 1939,<br />

Terashita 1985<br />

Galeola septentrionalis Armillaria jezoensis sp nov. Cha and Igarashi 1996<br />

Galeola sesamoides Fomes sp. New Zealand Campbell 1964<br />

Gastrodia cunninghamii Armillaria mellea Campbell 1962<br />

Gastrodia elata Armillaria mellea China Kusano 1911<br />

Gastrodia elata Armillariella mellea China Lan et al. 1994<br />

Gastrodia elata Mycena osmundicola New Zealand Lan et al. 1996<br />

Gastrodia minor brown basidiomycete w/clamps New Zealand Campbell 1962<br />

Gastrodia sesamoides Fomes mastoporus Campbell 1964<br />

Wullschlaegelia calcarata Acremonium, Colletotrichum, Curvularia,<br />

Puerto Rico J.T. Otero (unpublished)<br />

Cylindrocladium, Gliocladium, Paecilomyces,<br />

Penicillium, Trichoderma, Xylaria


162 P. Bayman, J.T. Otero<br />

Table 9.1. (continued)<br />

Orchids Fungi Location Reference<br />

Epiphytic & epilithic orchids<br />

Campylocentrum micranthum Calonectria kyotensis, Phomopsis cf orchidophila Costa Rica Richardson 1993<br />

Catasetum maculatum Acrogenospora sp., Codinaea parva, Colletotrichum crassipes, Costa Rica Richardson 1993<br />

Epicoccum andropogonis, Glomerella cingulata, Hadrotrichum sp.,<br />

Lasiodiplodia theobromae<br />

Catasetum maculatum Troposporella sp. Costa Rica Richardson<br />

and Currah 1995<br />

Dichaea standleyi Hadrotrichum sp., Nectria alata, Phomopsis cf. orchidophila Costa Rica Richardson 1993<br />

Dichaea trulla Colletotrichum crassipes Costa Rica Richardson 1993<br />

Dimerandra emarginata Epicoccum andropogonis, Hadrotrichum sp. Costa Rica Richardson 1993<br />

Dryadella pusiola Chaetomium homopilatum Costa Rica Richardson 1993<br />

Encyclia fragrans Alternaria alternata, Colletotrichum crassipes, Dactylaria sp., Costa Rica Richardson 1993<br />

Epicoccum andropogonis, Glomerella cingulata, Hadrotrichum sp.,<br />

Lasiodiplodia theobromae, Nodulisporum sp.,<br />

Pseudallescheriaboydii<br />

Epidendrum difforme Lasiodiplodia theobromae, Pithomyces maydicus Costa Rica Richardson 1993<br />

Epidendrum difforme Troposporella sp. Costa Rica Richardson and Currah 1995<br />

Epidendrum isomerum Nodulisporum sp. Costa Rica Richardson 1993<br />

Epidendrum nocturnum Epicoccum andropogonis Costa Rica Richardson 1993<br />

Epidendrum octomerioides Lasiodiplodia theobromae, Nectria ochroleuca,<br />

Costa Rica Richardson 1993<br />

Pestalotiopsis papposa, Phomopsis cf. orchidophila<br />

Epidendrum porpax Fusarium oxysporum, Guignardia sp. Colombia Dreyfuss and Petrini 1984<br />

Epidendrum schlechterianum Nectria haematococca, Periconiella sp., Pestalotiopsis papposa, Costa Rica Richardson 1993<br />

Xylaria sp.


9 Microbial Endophytes of Orchid Roots 163<br />

Table 9.1. (continued)<br />

Orchids Fungi Location Reference<br />

Epidendrum stangeanum Fusarium oxysporum, Hadrotrichum sp., Nodulisporium sp., Costa Rica Richardson 1993<br />

Pithomyces maydicus<br />

Epidendrum stangeanum Troposporella sp. Costa Rica Richardson and Currah 1995<br />

Epidendrum spp. Ascochyta sp., Colletotrichum gloeosporioides, Colletotrichum sp., Colombia/ Dreyfuss and Petrini 1984<br />

Cryptocline sp., Lasiodiplodia theobromae, Pestalotia cf. heterocornis, Brazil<br />

Phaeoseptoria cf. vermiformis, Phoma sp., Xylaria sp.<br />

Gongora unicolor Epicoccum andropogons, Hypoxylon cf. unitum, Lasmeniella sp., Costa Rica Richardson 1993<br />

Nodulisporium sp., Pestalotia poppola, Xylaria sp.<br />

Hexisea imbricata Arthrinium sp., Hadrotrichum sp., Pestalotiopsis aquatica Costa Rica Richardson 1993<br />

Jacquinella globosa Colletotrichum crassipes Costa Rica Richardson 1993<br />

Lepanthes caritensis Penicillium, Trichoderma, Xylaria corniformis Puerto Rico Tremblay et al. 1998<br />

Lepanthes rupestris Acremonium, Colletotrichum, Fusarium, Guignardia, Humicola, Puerto Rico Bayman et al. 2002<br />

Pestalotia, Phomposis, Trichoderma, Xylaria<br />

Lepanthes spp. Aspergillus, Colletotrichum, Penicillium, Pestalotia,<br />

Puerto Rico Bayman et al. 1997<br />

Xylaria arbuscula, X. corniformis, X. cf. cubensis,<br />

X. cf. curta multiplex, X. obovata, X. polymorpha, Xylaria sp.<br />

Maxillaria confusa Arthrinium sp., Malbranchea sp. Costa Rica Richardson 1993<br />

Maxillaria endresii Drechslera ellisii, Pestalotiopsis papposa Costa Rica Richardson 1993<br />

Maxillaria neglecta Chaetomium subspirale, Colletotrichum crassipes sp.,<br />

Costa Rica Richardson 1993<br />

Drechslera australensis, Glomerella cingulata, Hadrotrichum sp.,<br />

Humicola sp., Nectria haematococca, N. ochroleuca,<br />

Phomopsis cf. orchidophila, Xylaria sp.<br />

Maxillaria nicaraguensis Pestalotiopsis gracilis Costa Rica Richardson 1993<br />

Maxillaria uncata Cryptosporiopsis sp., Hadrotrichum sp. Costa Rica Richardson 1993<br />

Maxillaria xylobiflora Epicoccum nigrum Costa Rica Richardson 1993<br />

Maxillaria sp. Colletrotrichum crassipes, Hadrotrichum sp., Xylaria sp. Costa Rica Richardson 1993


164 P. Bayman, J.T. Otero<br />

Table 9.1. (continued)<br />

Orchids Fungi Location Reference<br />

Maxillaria sp. Acremonium strictum, Fusarium sambucinum, Tubercularia, Colombia, Dreyfuss and Petrini 1984<br />

Xylaria sp., Pestalotia cf. heterocornis, Phomopsis sp.<br />

Brazil<br />

Myoxanthus scandens Colletotrichum crassipes Costa Rica Richardson 1993<br />

Nidema boothii Dactylaria sp., Epicoccum andropogonis, Lasiodiplodia theobromae, Costa Rica Richardson and Currah 1995<br />

Leptosphaerulina australis, Pestalotiopsis papposa, Phomopsis cf.<br />

orchidophila<br />

Nidema boothii Troposporella sp. Costa Rica Richardson et al. 1993<br />

Octomeria sp. Melanotus alpiniae Costa Rica Richardson 1993<br />

Oncidium stenotis Chaetomium subspirale, Colletotrichum crassipes, Humicola sp., Costa Rica Richardson 1993<br />

Nigrospora sphaerica<br />

Pleurothallis corniculata Cryptosporiopsis sp., Pestalotiopsis papposa Costa Rica Richardson 1993<br />

Pleurothallis guanacastensis Hypoxylon cf. unitum, Lasiodiplodia theobromae,<br />

Costa Rica Richardson 1993<br />

Pithomyces maydicus, Xylaria sp.<br />

Pleurothallis pantasmi Hadrotrichum sp., Humicola sp., Nectria peziza Costa Rica Richardson 1993<br />

Pleurothallis periodica Chaetomium subspirale, Cladosporium cladosporioides,<br />

Costa Rica Richardson 1993<br />

Geotrichopsis sp., Hadrotrichum sp., Xylaria sp.<br />

Pleurothallis phyllocardioides Lasiodiplodia theobromae, Nectria haematococca Costa Rica Richardson 1993<br />

Pleurothallis uncinata Nectria haematococca, Pithomyces maydicus Costa Rica Richardson 1993<br />

Pleurothallis verecunda Epicoccum andropogonis Costa Rica Richardson 1993<br />

Pleurothallis sp. Chaetomium funicola, Pyrenochaeta cf. rubi-idaei,<br />

Costa Rica Richardson 1993<br />

Ramichloridium cf. subulatum<br />

Pleurothallis sp. Cryptocline sp., Xylaria sp. Colombia Dreyfuss and Petrini 1984<br />

Polystachya foliosa Hadrotrichum sp., Xylaria sp. Costa Rica Richardson 1993<br />

Psychilis kraenzlinii Colletotrichum, Curvularia, Mucor, Paecilomyces, Pestalotia, Puerto Rico J.T. Otero (unpublished)<br />

Xylaria


9 Microbial Endophytes of Orchid Roots 165<br />

Table 9.1. (continued)<br />

Orchids Fungi Location Reference<br />

Psychilis krugii Colletotrichum, Curvularia, Mucor, Paecilomyces, Pestalotia, Puerto Rico J.T. Otero (unpublished)<br />

Xylaria<br />

Rodriguezia compacta Colletotrichum crassipes, Epicoccum andropogonis,<br />

Costa Rica Richardson 1993<br />

Hadrotrichum sp., Leptosphaerulina australis,<br />

Pestalotiopsis aquatica<br />

Scaphyglottis cf. prolifera Nectria ochroleuca Costa Rica Richardson 1993<br />

Scaphyglottis gracilis Nectria haematococca, N. ochroleuca Costa Rica Richardson 1993<br />

Scaphyglottis minutiflora Arthrinium sp. Costa Rica Richardson 1993<br />

Sobralia cf. mucronata Arthrinium sp., Lasiodiplodia theobromae Costa Rica Richardson 1993<br />

Sobralia mucronata Pseudallescheria boydii Costa Rica Richardson 1993<br />

Sobralia powellii Xylaria sp. Costa Rica Richardson 1993<br />

Sobralia sp. Epicoccum nigrum, Glomerella cingulata,<br />

Costa Rica Richardson 1993<br />

Lasiodiplodia theobromae, Nectria haematococca<br />

Stelis endresii Curvularia cymbopogonis, Nectria haematococca,<br />

Costa Rica Richardson 1993<br />

Phomopsis cf. orchidophila<br />

Stelis sp. Alternaria alternata, Arthrinium sp., Chloridium virescens, Costa Rica Richardson 1993<br />

Colletotrichum crassipes, Cryptosporiopsis sp.,<br />

Epicoccum andopogensis, Hadrotrichum sp., Hypoxylon cf. unitum,<br />

Nectria alata, Nectria ochroleuca, N. radicicola,<br />

Neoplaconema napelli, Pithomyces maydicus, Xylaria spp.<br />

Trichosalpinx blaisdellii Lasiodiplodia theobromae Costa Rica Richardson 1993<br />

Trichosalpinx orbicularis Chaetosticta cf. perforata, Colletotrichum crassipes, Nectria alata Costa Rica Richardson 1993<br />

Trichosalpinx sp. Hadrotrichum sp. Costa Rica Richardson 1993<br />

Trigonidium egertonianum Chaetosticta cf. perforata, Epicoccum andopogensis Costa Rica Richardson 1993<br />

Trigonidium riopalaquense Chaetomium aureum, Colletotrichum acutatum, C. crassipes, Costa Rica Richardson 1993<br />

Nectria haematococca, Nodulisporium sp., Periconiella sp.


166 P. Bayman, J.T. Otero<br />

Table 9.1. (continued)<br />

Orchids Fungi Location Reference<br />

Unidentified Acremonium pteridii, Anthostomella aracearum, Ascochyta sp., French Petrini and Dreyfuss 1981<br />

Aureobasidium caulivorum, Chaetosphaeria endophytica, Guayana<br />

Colletotrichum spp., Coniothyrium sp., Cryptocline spp.,<br />

Cryptosporiopsis sp., Curvularia pallescens, Cytogloeum sp.,<br />

Fusarium oxysporum, Gelatinosporium spp., Gliocladium roseum,<br />

Glomerella cingulata, Kaskaskia sp., Lasiodiplodia theobromae,<br />

Melanconium sp., Microascus cinereus, Microcyclus sp.,<br />

Nodulisporium gregarium, Nodulisporium spp., Pestalotia adusta,<br />

Phialaspora sp., Phoma sp., Phomatospora berkeleyi,<br />

Phomopsis orchidophila, Phomopsis sp., Phyllosticta capitalensis,<br />

P. colocasiicola, Ramichloridium apiculatum, Verticillium lecanii<br />

aThe following fungal genera are presumed to be mycorrhizal and are not included in this table: Ceratobasidium, Oliveonia Sebacina, Serendipita,<br />

Thanatephorus, Tulanella and Ypsilonidium (teleomorphs); Ceratobasidium (anamorphs) (Roberts 1999). These fungi can be found in tables<br />

published by Currah et al. (1997), Rasmussen (2002) and Taylor et al. (2002)<br />

bAlso called Mycelium radicis atrovirens (MRA)<br />

c This species can be either myco-heterotrophic or photosynthetic


9 Microbial Endophytes of Orchid Roots 167<br />

mycorrhizae in some plants (Fernando and Currah 1996; Jumpponen 2001),<br />

but their role in orchids is still unclear (Rasmussen 2002).<br />

One of the most ubiquitous and interesting groups of endophytes is<br />

Fusarium and its teleomorphs. Fusarium moniliforme (= F. verticilloides)<br />

wasisolatedfromleavesandrootsofCypripedium reginae (Peschke and<br />

Volz 1978). Inoculation of spores induced disease symptoms in hybrid<br />

orchids, suggesting that the fungus was a potential pathogen (Peschke and<br />

Volz 1978). This fungus is a ubiquitous endophyte and pathogen of all<br />

organs of various hosts, including maize (Leslie et al. 1990; Kuldau and<br />

Yates 2000; see Chap. 8 by Bacon and Yates).<br />

On the other hand, the ability of Fusarium to stimulate orchid seed<br />

germination has been known for 100 years (Bernard 1909). An unidentified<br />

Fusarium strain isolated from a germinating seed of Cypripedium reginae<br />

induced germination of C. reginae seeds in vitro (Vujanovic et al. 2000).<br />

Although this is functional rather than ecological specificity (Masuhara<br />

and Katsuya 1994), it raises the question of whether a single Fusarium<br />

isolate can be endophytic, pathogenic and mycorrhizal under different<br />

circumstances.<br />

9.8<br />

Endophytic Fungi in Roots of Myco-Heterotrophic Orchids<br />

Non-photosynthetic (or more precisely, myco-heterotrophic) orchids have<br />

been extensively studied because of their interesting relationships with<br />

fungi (Leake 1994). Unable to assimilate their own carbon, these orchids<br />

are parasitic on fungi. Several myco-heterotrophic orchids are very specific<br />

for certain fungi, which is interesting because orchid mycorrhizal relationships<br />

are generally considered to be non-specific (Taylor et al. 2002). In<br />

most cases the myco-heterotrophic orchids are parasitizing a mycorrhizal<br />

partner of a nearby photosynthetic plant, which means that they are indirectly<br />

parasitizing the plant as well. However, the amount of carbon taken<br />

by the orchid is probably insignificant to the plant host (McKendrick et<br />

al. 2000; Sanders 2003). An excellent review of mycorrhizal specificity in<br />

myco-heterotrophic plants is available (Taylor et al. 2002).<br />

Fungal DNA has been amplified directly from roots or pelotons of mycoheterotrophic<br />

orchids using fungal-specific (or in some cases, basidiomycete-specific)<br />

primers. This approach has been used on several orchids<br />

in North America: Cephalanthera (Taylor and Bruns 1997), Corallorhiza<br />

spp. (Taylor and Bruns 1997, 1999) and Hexalectris (Taylor et al. 2003). It<br />

has also been used on Dactylorhiza in Denmark (Kristiansen et al. 2001)<br />

and Neottia in the United Kingdom, Germany (McKendrick et al. 2002)<br />

and in France (Selosse et al. 2002). In all these plants, the only fungi


168 P. Bayman, J.T. Otero<br />

amplified from peletons belonged to taxa that are considered ectomycorrhizal.<br />

No other endophytic fungi were reported, which may suggest that<br />

secondary colonization of pelotons by non-mycorrhizal endophytic fungi<br />

is uncommon. No such studies have been done with the aim of identifying<br />

non-mycorrhizal fungi in orchids, e.g., using ascomycete-specific PCR<br />

primers.<br />

Myco-heterotrophic orchids tend to associate with ectomycorrhizal fungi<br />

rather than the Rhizoctonia-like fungi typical of orchids, presumably because<br />

ectomycorrhizal fungi are more reliable suppliers of photosynthate.<br />

In other cases, the presumed mycorrhizal fungi are basidiomycetes that are<br />

wood decomposers not known to form ectomycorrhizae, e.g., Armillaria,<br />

Fomes, Ganoderma and Phellinus sp. (for references, see Rasmussen 2002;<br />

Taylor et al. 2002). However, most such studies have not demonstrated<br />

functional relationships between the fungus and the myco-heterotrophic<br />

orchid, so this is another area where the boundary between mycorrhizal<br />

fungi and endophytes is unclear.<br />

Endophytes have been reported from only one tropical myco-heterotrophic<br />

orchid, Wullschlaegelia. Forty-one morphospecies of endophytes<br />

were isolated from plants of W. calcarata in Puerto Rico (J.T. Otero, unpublished;<br />

Table 9.1). Common endophytes included Xylaria, Trichoderma, Colletotrichum<br />

and various dematiaceous hyphomycetes. These genera were<br />

not randomly distributed among sites (χ 2 = 31.84, df = 17, P = 0.004).<br />

A morphospecies accumulation curve suggested that most of the culturableendophyticfungiintherootswereisolated(Fig.9.1).Itislikelythat<br />

temperate myco-heterotrophic orchids contain a mycoflora as diverse as<br />

W. calcarata, but the peloton isolation technique used in many of the above<br />

studies excluded most of the non-mycorrhizal endophytes.<br />

Fig.9.1. Species accumulation curve for endophytic fungi isolated from roots of<br />

Wullschlaegelia calcarata, a myco-hetrotrophic orchid. A total of 32 plants were collected<br />

from eight populations in El Verde, Puerto Rico. Morphospecies were identified by morphology<br />

in culture; the identified fungi are listed in Table 9.1


9 Microbial Endophytes of Orchid Roots 169<br />

9.9<br />

Endophytic Fungi in Roots of Epiphytic<br />

and Lithophytic Orchids<br />

Reports on non-mycorrhizal endophytes in roots of epiphytic and lithophytic<br />

orchids have mostly come from the neotropics. These reports have<br />

focused on identifying the fungi rather than exploring their relationship<br />

with the orchids. Descriptions of some common endophytes that have been<br />

published will facilitate further studies (Richardson et al. 1993; Richardson<br />

and Currah 1995; Currah et al. 1997).<br />

South America Three taxa of Ascomycetes, 5 of Hyphomycetes and 13 of<br />

Coelomycetes were isolated from epiphytic orchid roots in French Guiana;<br />

many of these fungi were also isolated from roots of aroids and bromeliads,<br />

suggesting that the fungi are generalists (Table 9.1; Petrini and Dreyfuss<br />

1981). Some of the same fungi were also found in orchid roots from the<br />

Colombian Amazon (Dreyfuss and Petrini 1984).<br />

Costa Rica The most extensive sampling of epiphytic orchid roots for endophytes<br />

was done in La Selva, Costa Rica (Richardson et al. 1993; Richardson<br />

and Currah 1995). Of 59 species of epiphytic orchids sampled in La<br />

Selva, Costa Rica, mycorrhizal pelotons were observed in the roots of 23<br />

(= 39%) (Richardson et al. 1993), a fairly low infection frequency that<br />

agrees with other studies. Basidiomycetes comprised only 3% of the fungi<br />

isolated, suggesting that mycorrhizal fungi were much less common than<br />

non-mycorrhizal endophytes or pathogens. Hadrotrichum, Colletotrichum,<br />

Epicoccum, Lasiodiplodia and Phomopsis were the most common genera<br />

of deuteromycetes and ascomycetes (Richardson and Currah 1995). Relative<br />

proportions of ascomycetes, hyphomycetes and coelomycetes were<br />

comparable to those reported by Petrini and Dreyfuss (1981).<br />

Puerto Rico We have sampled endophytic fungi from roots of various epiphytic<br />

orchids in Puerto Rico (Table 9.1). The most common endophytic<br />

fungi have been fairly consistent from study to study. A variety of endophytic<br />

fungi were isolated from nine species of epiphytic orchids in Puerto<br />

Rico, including Xylaria, Pestalotia and Colletotrichum (Otero et al. 2002).<br />

Rhizoctonia-like fungi were isolated at lower frequency, from about 20% of<br />

the samples.<br />

Fifty-five fungi (in 26 morphospecies) were isolated from roots of Tolumnia,<br />

from both juvenile and adult plants (J.T. Otero, unpublished).<br />

Thirty-one of these strains (in 13 morphospecies) were tested for potential<br />

mycorrhizal activity with T. variegata seeds. Thirteen strains (in 6 morphospecies)<br />

had a positive effect on seed germination in vitro, all of which


170 P. Bayman, J.T. Otero<br />

were Rhizoctonia-like fungi; others were parasitic on seeds. These data<br />

suggest Rhizoctonia are the principal mycorrhizal fungi of T. variegata, but<br />

thesampledidnotincludeFusarium (see Sect. 9.9).<br />

A comparison of two sympatric populations of Psychilis and P. krugii<br />

found 26 morphospecies of endophytic fungi (Table 9.1; J.T. Otero, unpublished).<br />

There was no apparent difference between the fungal communities<br />

of the two orchids. The dominant species were Xylaria spp.; Rhizoctonialikefungiweremuchlesscommonthanintheorchidscitedabove,andfew<br />

pelotons were seen in root sections.<br />

Endophytic fungi were isolated from roots and leaves of six species of<br />

Lepanthes in Puerto Rico (Bayman et al. 1997). Xylaria and Rhizoctonia<br />

were the most common genera isolated from roots and leaves. At least nine<br />

species of Xylaria wereisolated,uptofourspeciesoccurringinasingle<br />

plant. Frequency of Xylaria species and Rhizoctonia did not differ significantly<br />

between roots and leaves, which is perhaps not surprising given that<br />

roots of epiphytes are exposed to light and air. Frequency of both Rhizoctonia<br />

and Xylaria differed significantly among Lepanthes species. However,<br />

there was also significant variation among different roots of a single plant,<br />

which means that studies that wish to compare levels of infection should<br />

sample intensively. In another study of L. rupestris,endophyticfungiwere<br />

isolated from roots of lithophytic plants (Bayman et al. 2002). The most<br />

common genera were Guignardia (isolated from 22% of root pieces), Colletotrichum<br />

(10%) and Xylaria (7%). Rhizoctonia-like fungi (which include<br />

the presumed mycorrhizal fungi) were isolated at much lower frequencies<br />

(3%).<br />

Is distribution of fungi a limiting factor in distribution of orchids? Tremblay<br />

et al. (1998) isolated fungi from roots of Lepanthes caritensis,anorchid<br />

species whose distribution is limited to a few trees along a single river. Fungi<br />

isolated from orchid roots were compared to fungi isolated from bark of<br />

host trees, and of conspecific trees without orchids (Tremblay et al. 1998).<br />

The most common fungi in L. caritensis roots were Xylaria spp, particularly<br />

X. corniformis. Penicillium, Trichoderma and Rhizoctonia were also<br />

isolated from roots, and were the most common fungi isolated from tree<br />

bark. However, Xylaria sp. were not common on tree bark, suggesting that<br />

they had a particular affinity for Lepanthes roots. The number of other,<br />

unidentified fungi was significantly higher on bark of trees without orchids<br />

than on trees with orchids. This may suggest that the presence of certain<br />

fungi inhibits the establishment of orchids.<br />

In general, there is little evidence that non-mycorrhizal endophytes in<br />

orchids are specific to orchids. The most common groups of orchid endophytes<br />

(Table 9.1) are fungi that are ubiquitous in soil and as endophytes<br />

of other plants. There are marked differences between terrestrial and epiphytic<br />

orchids: in terrestrials, Phialocephala is the most frequently isolated


9 Microbial Endophytes of Orchid Roots 171<br />

group; in epiphytes, Hadrotrichum, Epicoccum, Lasiodiplodia, Xylaria and<br />

Pestalotiopsis are most frequent. However, these differences reflect differences<br />

in temperate vs. tropical mycofloras and are not particular to orchidassociated<br />

fungi. Several new species have been described from orchid<br />

endophytes (e.g., Mycena sp. nov. (Fan et al. 1996), Armillaria jezoensis sp.<br />

nov. (Cha and Igarashi 1996), Leptodontidium (Currah et al. 1987)), but in<br />

most cases their distribution is not sufficiently known to claim a special<br />

affinity for orchids.<br />

9.10<br />

Endophytic Fungi in Epiphytic Orchid Roots:<br />

Importance to Plant Hosts<br />

There are two reasons to believe that the presence of endophytes could<br />

affect mycorrhizal fungi, and vice versa. First, orchids may produce phytoalexins<br />

such as orchinol when challenged by a fungus; these phytoalexins<br />

may then limit the ability of other fungi to colonize the plant – a type of<br />

induced resistance. These phytoalexins inhibit growth of a broad range of<br />

fungi, though bacteria are less susceptible (Gäumann et al. 1960). According<br />

to Rasmussen (1995), “...all underground parts of terrestrial orchids<br />

must either accommodate the endophyte (i.e., mycorrhizal fungus) or actively<br />

reject it.” Second, endophytes and mycorrhizal fungi could compete<br />

for nutrients in the root, or could actively inhibit each other by production<br />

of secondary metabolites. Nutrient translocation from mycorrhizal fungi<br />

to orchids has been demonstrated repeatedly (see Rasmussen 1995; Bidartondo<br />

et al. 2004), but it is unknown how non-mycorrhizal endophytes<br />

might affect this transfer.<br />

Several studies have asked whether the presence of endophytic fungi<br />

was associated with the presence of other endophytes or of mycorrhizal<br />

fungi. Pieces of Lepanthes roots colonized by Colletotrichum had a significantly<br />

lower rate of infection with Xylaria thanwouldbeexpectedfrom<br />

the frequency of each genus alone (Bayman et al. 2002). Presence of Colletotrichum<br />

also showed a significant, negative correlation with Guignardia.<br />

This suggests there may be competition or antagonism between these<br />

genera;alternatively,theircolonizationorgrowthcouldbefavoredbydifferent<br />

environmental conditions. Also, mycorrhizal fungi in Vanilla often<br />

occur together with other, presumably pathogenic, fungi (Alconcero 1969;<br />

Porras-Alfaro and Bayman 2003). Roots of Vanilla plants in Puerto Rico<br />

were often colonized by both the pathogen Fusarium oxysporum and by<br />

Rhizoctonia solani, which was both mycorrhizal and pathogenic to Vanilla<br />

(Alconcero 1969).


172 P. Bayman, J.T. Otero<br />

Fungicides were applied to plants of L. rupestris to see if the costs of<br />

harboring fungi outweighed the benefits (Bayman et al. 2002). Propiconazole<br />

significantly reduced the number of total fungal colonies isolated from<br />

roots (but not from leaves), increased plant mortality, and increased loss<br />

of leaves, as compared to control plants. Benomyl significantly reduced<br />

the number of fungi isolated from leaves (but not from roots) and decreased<br />

plant mortality, but did not significantly effect plant growth. These<br />

data suggest that fungi have both positive and negative effects on growth<br />

and survival of orchid plants in the field: benomyl, which affects mainly<br />

ascomycete fungi, may have reduced pathogens and endophytes, not harming<br />

plants, whereas propiconazole, which also affects basidiomycetes, may<br />

have reduced mycorrhizal fungi as well. However, these results are very<br />

difficult to interpret: a single plant may have simultaneous infections by<br />

endophytes, mycorrhizal fungi and pathogens, and the positive effects of<br />

one may mask the negative effects of another.<br />

The interaction between plants and endophytic fungi may be viewed as<br />

a ‘balanced antagonism’ (Schulz et al. 1999). The plant produces secondary<br />

metabolites that are sufficiently toxic to restrict the growth of the endophyte<br />

without being able to kill it; the endophyte in turn produces enzymes and<br />

metabolites that allow it to colonize the plant but are not sufficient to cause<br />

pathogenicity. This balance becomes still more complex if the endophyte<br />

and its activities also affect other fungi, either endophytic or mycorrhizal,<br />

as these data from Lepanthes suggest.<br />

9.11<br />

Conclusions<br />

About 90% of the cells that comprise a human body are microbial – we are<br />

walking communities (Hamilton 1999). Most of these microorganisms are<br />

poorly understood: pathogenic microorganisms are intensively studied,<br />

but much less is known about non-pathogenic, commensal microorganisms,<br />

which are more common than pathogens. Their importance becomes<br />

obvious when an antibiotic designed to kill a pathogen affects the community<br />

of commensals as well, causing a secondary condition – for example,<br />

vaginal yeast infections in women who are taking antibiotics for bladder infections.<br />

Cross-talk between microorganisms and the host may have other<br />

roles as well; for example, it may be necessary for the proper development<br />

of the immune system (Hooper et al. 1998). Even well-studied pathogens<br />

may have complex ecological roles: Helicobacter was considered a serious<br />

pathogen, but recent studies suggest that it is a commensal that sometimes<br />

turns pathogenic (Pütsep et al. 1999). Our ignorance of the non-pathogenic<br />

microflora is a limiting factor in medicine.


9 Microbial Endophytes of Orchid Roots 173<br />

The same situation exists with orchid root endophytes. Interest in orchid<br />

mycorrhizal fungi has overshadowed work on other orchid endophytes,<br />

which may have important interactions with mycorrhizal fungi and with<br />

the orchids themselves. In some cases the distinction between endophytes<br />

and mycorrhizal fungi is unclear, and it is likely that, as with H. pylori<br />

in humans, a single organism can behave as an endophyte, mycorrhizal<br />

symbiont or pathogen, depending on the environment and the health of<br />

the host plant. The study of orchid endophytes is not much more advanced<br />

than the study of orchid mycorrhizae was in 1900 – interesting observations,<br />

intriguing ideas, but little information about functional significance.<br />

<strong>References</strong><br />

Alconero R (1969) Mycorrhizal synthesis and pathology of Rhizoctonia solani in Vanilla<br />

orchid roots. Phytopathology 59:526–530<br />

Andersen TF, Rasmussen HN (1996) The mycorrhizal species of Rhizoctonia. In: Sneh B,<br />

Jabaji-Hare S, Neate S, Dijst G (eds) Rhizoctonia species: taxonomy, molecular biology,<br />

ecology, pathology and disease control. Kluwer, Dordrecht, pp 379–390<br />

Arditti J, Ernst R, Yam TW, Glabe C (1990) The contribution of orchid mycorrhizal fungi to<br />

seed germination: a speculative review. Lindleyana 5:249–255<br />

Arnold AE, Maynard Z, Gilbert G, Coley PD, Kursar TA (2000) Are tropical fungal endophytes<br />

hyperdiverse? Ecol Lett 3:267–274<br />

Arnold AE, Maynard Z, Gilbert GS (2001) Fungal endophytes in dicotyledonous neotropical<br />

trees: patterns of abundance and diversity. Mycol Res 105:1502–1507<br />

Bayman P, Lebrón LL, Tremblay RL, Lodge DJ (1997) Fungal endophytes in roots and leaves<br />

of Lepanthes (Orchidaceae). New Phytol 135:143–149<br />

Bayman P, González EJ, Fumero JJ, Tremblay RL (2002) Are fungi necessary? How fungicides<br />

affect growth and survival of the orchid Lepanthes rupestris in the field. J Ecol 90:1002–<br />

1008<br />

Bernard N (1909) L’évolution dans la symbiose. Le orchidées et leurs champignons commensaux.<br />

Ann Sci Nat Bot 9:1–196<br />

Bidartondo MI, Burghardt B, Gebauer G, Bruns TD, Read DJ (2004) Changing partners in<br />

the dark: isotopic and molecular evidence of ectomycorrhizal liaisons between forest<br />

orchids and trees. Proc R Soc Lond B 271:1799–1806<br />

Brighigna L,Montaini P,Favilli F,Carabez-Trejo A(1992) Role of the nitrogen-fixingbacterial<br />

microflora in the epiphytism of Tillandsia (Bromeliaceae). Am J Bot 79:723–727<br />

Bronstein JL, Wilson WG, Morris WF (2003) Ecological dynamics of mutualist/antagonist<br />

communities. Am Nat 162:S24–S39<br />

Bruns TD, Szaro TM, Gardes M, Cullings KW, Pan JJ, Taylor DL, Horton TR, Kretzer A,<br />

Garbelotto M, Li Y (1998) A sequence database for the identification of ectomycorrhizal<br />

basidiomycetes by phylogenetic analysis. Mol Ecol 7:257–272<br />

Burgeff H (1959) Mycorrhiza of orchids. In: Withner CL (ed) The orchids: a scientific survey.<br />

Ronald, New York, pp 361–395<br />

Campbell EO (1962) The mycorrhiza of Gastrodia cunninghamii Hook.TransRSocNZ<br />

1:289–296<br />

Campbell EO (1964) The fungal association in a colony of Gastrodia sesamoides. R. Br Trans<br />

R Soc N Z 2:237–246


174 P. Bayman, J.T. Otero<br />

Campbell EO (1970a) Morphology of the fungal association in three species of Corallorhiza<br />

in Michigan. Mich Bot 9:108–113<br />

Campbell EO (1970b) The fungal association of Yoania australis. Trans R Soc N Z Biol Sci<br />

12:5–12<br />

Carling DE, Pope EJ, Brainard KA, Carter DA (1999) Characterization of mycorrhizal isolate<br />

of Rhizoctonia solani from orchid, including AG-12, a new anastomosis group.<br />

Phytopathology 89:942–946<br />

Cha JY, Igarashi T (1996) Armillaria jezoensis, anewsymbiontofGaleola septentrionalis<br />

(Orchidaceae) in Hokkaido. Mycoscience 37:21–24<br />

Chanway CP (1995) Endophytes: they’re not just fungi! Can J Bot 74:321–233<br />

Currah RS (1991) Taxonomic and developmental aspects of the fungal endophytes of terrestrial<br />

orchid mycorrhizae. Lindleyana 6:211–213<br />

Currah RS, Sigler L, Hambleton S (1987) New records and taxa of fungi from the mycorrhizae<br />

of terrestrial orchids of Alberta. Can J Bot 65:2473–2482<br />

Currah RS, Smreciu EA, Hambleton S (1990) Mycorrhizae and mycorrhizal fungi of<br />

boreal species of Platanthera and Coeloglossum (Orchidaceae). Can J Bot 68:1171–<br />

1181<br />

Currah RS, Zelmer CD, Hambleton S, Richardson KA (1997) Fungi from orchid mycorrhizas.<br />

In: Arditti J, Pridgeon A (eds) Orchid biology: reviews and perspectives, VII. Kluwer,<br />

Lancaster, pp 117–170<br />

Darwin C (1887) The various contrivances by which orchids are fertilized by insects, 2nd<br />

edn. William Clowes and Sons, London<br />

Dressler RL (1990) The orchids: natural history and classification. Harvard University Press,<br />

Cambridge MA<br />

Dreyfuss M, Petrini O (1984) Further investigations on the occurrence and distribution of<br />

endophytic fungi in tropical plants. Bot Helv 94:33–40<br />

Fan L, Guo S, Cao W, Xiao P, Xu J, Fan L, Guo SX, Cao WQ, Xiao PG, Xu JT (1996)<br />

Isolation, culture, identification and biological activity of Mycena orchidicola sp. nov.<br />

in Cymbidium sinense (Orchidaceae). Acta Mycol Sin 15:251–255<br />

Fernando AA, Currah RS (1995) Leptodontidium orchidicola (Mycelium radicis atrovirens<br />

complex): aspects of its conidiogenesis and ecology. Mycotaxon 54:287–294<br />

Fernando AA, Currah RS (1996) A comparative study of the effects of the root endophytes<br />

Leptodontidium orchidicola and Phialocephala fortinii (Fungi Imperfecti) on the growth<br />

of some subalpine plants in culture. Can J Bot 74:1071–1078<br />

Fröhlich J, Hyde KD (1999) Biodiversity of palm fungi in the tropics: are global fungal<br />

diversity estimates realistic? Biodivers Conserv 8:977–1004<br />

Gamboa MA, Bayman P (2001) Communities of endophytic fungi in leaves of a tropical<br />

timber tree (Guarea guidonia: Meliaceae). Biotropica 33:352–360<br />

Gäumann E, Nüesch J, Rimpau RH (1960) Weitere Untersuchungen über die chemischen<br />

Abwehrreaktionen der Orchideen. Phytopathol Z 38:274–308<br />

Goh CJ, Sim AA, Lim G (1992) Mycorrhizal associations in some tropical orchids. Lindleyana<br />

7:13–17<br />

Hadley G, Williamson B (1972) Features of mycorrhizal infection in some Malayan orchids.<br />

New Phytol 71:1111–1118<br />

Hadley G, Ong SH (1978) Nutritional requirements of orchid endophytes. New Phytol<br />

81:561–569<br />

Hadley G, Arditti M, Arditti J (1987) Orchid diseases: a compendium. In: Arditti J (ed)<br />

Orchid biology: reviews and perspectives, vol IV. Cornell University Press, Ithaca, NY,<br />

pp 261–325<br />

Hallmann J, Quadt-Hallmann, Mahaffee WF, Kloepper JW (1997) Bacterial endophytes in<br />

agricultural crops. Can J Microbiol 43:895–914


9 Microbial Endophytes of Orchid Roots 175<br />

Hamada M (1939) Studien über die Mykorrhiza von Galeola septentrionalis Reichb. f. neuer<br />

Fall der Mykorrhiza-Bildung durch intraradicale Rhizomorpha. Jpn J Bot 10:151–211<br />

Hamada M, Nakamura SI (1963) Wurzelsymbiose von Galeola altissima Reichb. F., einer<br />

chlorophyllfreien Orchidee, mit dem holzzerstörenden Pilz Hymenochate crocicreas.<br />

Berk Et Br. Sci Rep Tohoku Univ Ser IV (Biol) 29:227–238<br />

Hamilton G (1999) Insider trading. New Scientist 6:42–46<br />

Hawksworth DL (1991) The fungal dimension of biodiversity: magnitude, significance and<br />

conservation. Mycol Res 95:641–655<br />

Hawksworth DL (2000) How many fungi are there? Mycol Res 104:4–5<br />

Hawksworth DL, Rossman AY (1997) Where are all the undescribed fungi? Phytopathology<br />

87:888–891<br />

Hooper LV, Bry L, Falk PG, Gordon JI (1998) Host-microbial symbiosis in the mammalian<br />

intestine: exploring an internal ecosystem. BioEssays 20:336–343<br />

Jonsson L, Nylund JE (1979) Favolaschia dybowskiana (Singer) Singer (Aphyllophorales),<br />

a new orchid mycorrhizal fungus from tropical Africa. New Phytol 83:121–128<br />

Jumpponen A (2001) Dark septate endophytes – are they mycorrhizal? Mycorrhiza<br />

11:207–211<br />

Kristiansen KA, Taylor DL, Kjoller R, Rasmussen HN, Rosendahl S (2001) Identification<br />

of mycorrhizal fungi from single pelotons of Dactylorhiza majalis (Orchidaceae) using<br />

single-strand conformation polymorphism and mitochondrial ribosomal large subunit<br />

DNA sequences. Mol Ecol 10:2089–2093<br />

Kuldau GA, Yates IE (2000) Evidence for Fusarium endophytes in cultivated and wild plants.<br />

In: Bacon CW, White JF (eds) Microbial endophytes, Dekker, New York, pp 85–117<br />

Kusano S (1911) Gastrodia elata and its symbiotic association with Armillaria mellea.JColl<br />

Agric Jpn 9:1–73<br />

Lacap DC, Hyde KD, Liew ECY (2003) An evaluation of the fungal “morphotype” concept<br />

based on ribosomal DNA sequences Fungal Divers 12:55–63<br />

Lan J, Xu JT, Li JS (1994) Study on symbiotic relation between Gastrodia elata and Armillariella<br />

mellea by autoradiography. Acta Mycol Sin 13:219–222<br />

Lan J, Xu JT, Li J (1996) Study on infecting process of Mycena osmundicola on Gastrodia<br />

elata by autoradiography. Acta Mycol Sin 13:219–222<br />

Leake JR (1994) Tansley Review No. 69: the biology of myco-heterotrophic (“saprophytic”)<br />

plants. New Phytol 127:171–216<br />

Leslie JF, Pearson CAS, Nelson PE, Tousson TA (1990) Fusarium spp. from corn, sorghum<br />

and soybean fields in the central and eastern United States. Phytopathology 80:343–350<br />

Lesica P, Antibus RK (1990) The occurrence of mycorrhizae in vascular epiphytes of two<br />

Costa Rican rain forests. Biotropica 22:250–258<br />

Masuhara G, Katsuya K (1994) In situ and in vitro specificity between Rhizoctonia spp. and<br />

Spiranthes sinensis (Persoon) Ames. var. amoena (M. Bieberstein) Hara (Orchidaceae).<br />

New Phytol 127:711–718<br />

McKendrick SL, Leake JR, Read DJ (2000) Symbiotic germination and development of mycoheterotrophic<br />

plants in nature: transfer of carbon from ectomycorrhizal Salix repens and<br />

Betula pendula to the orchid Corallorhiza trifida through shared hyphal connections.<br />

New Phytol 145:539–548<br />

McKendrick SL, Leake JR, Taylor DL, Read DJ (2002) Symbiotic germination and development<br />

of the myco-heterotrophic orchid Neottia nidus-avis in nature and its requirement<br />

for locally distributed Sebacina spp. New Phytol 154:233–247<br />

Otero JT, Ackerman JD, Bayman P (2002) Diversity and host specificity of mycorrhizal fungi<br />

from tropical orchids. Am J Bot 89:1852–1858<br />

Peschke HC, Volz PA (1978) Fusarium moniliforme Sheld association with species of orchids.<br />

Phytologia 40:347–356


176 P. Bayman, J.T. Otero<br />

Petrini O, Dreyfuss M (1981) Endophytische Pilze in epiphytischen Araceae, Bromeliaceae<br />

und Orchidaceae. Sydowia 34:135–148<br />

Porras-Alfaro A, Bayman P (2003) Mycorrhizal fungi of Vanilla: root colonization patterns<br />

and fungal identification. Lankesteriana 7:147–150<br />

Pütsep K, Brändén CI, Boman HG, Normark S (1999) Antibacterial peptide from H. pylori.<br />

Nature 398:671–672<br />

Ramsay RR, Dixon KW, Sivasithamparam K (1986) Patterns of infection and endophytes<br />

associated with Western Australian orchids. Lindleyana 1:203–214<br />

Rasmussen HN (1995) Terrestrial orchids from seed to mycotrophic plant. Cambridge<br />

University Press, Cambridge UK<br />

Rasmussen HN (2002) Recent developments in the study of orchid mycorrhiza. Plant Soil<br />

244:149–163<br />

Richardson KA (1993) Endophytic fungi from Costa Rican orchids. MSc Thesis, University<br />

of Alberta, Edmonton<br />

Richardson KA, Currah RS (1995) The fungal community associated with the roots of some<br />

rainforest epiphytes of Costa Rica. Selbyana 16:49–73<br />

Richardson KA, Currah RS, Hambleton S (1993) Basidiomycetous endophytes from the<br />

roots of neotropical epiphytic Orchidaceae. Lindleyana 8:127–137<br />

Rivas M, Warner J, Bermúdez M (1998) Presence of mycorrhizas in orchids of a neotropical<br />

botanical garden. Rev Biol Trop 46:211–216<br />

Roberts P (1999) Rhizonctonia-forming fungi. A taxonomic guide. The trustees of the Royal<br />

Botanical Gardens, Kew, London<br />

Salmia A (1988) Endomycorrhizal fungus in chlorophyll-free and green forms of the terrestrial<br />

orchid Epipactis helleborine. Karstenia 28:3–18<br />

Sanders IR (2003) Preference, specificity and cheating in the arbuscular mycorrhizal symbiosis.<br />

Trends Plant Sci 8:143–154<br />

Schulz B, Römmert AK, Dammann U, Aust HJ, Strack D (1999) The endophyte-host interaction:<br />

a balanced antagonism? Mycol Res 103:1275–1283<br />

Selosse M-A, Weiß M, Jany JL, Tillier A (2002) Communities and populations of sebacinoid<br />

basidiomycetes associated with the achlorophyllous orchid Neottia nidus-avis and<br />

neighboring tree ectomycorrhizae. Mol Ecol 11:1831–1844<br />

Selosse M-A, Faccio A, Scappaticci G, Bonfante P (2004) Chlorophyllous and achlorophyllous<br />

specimens of Epipactis microphylla (Neottieae, Orchidaceae) are associated with<br />

ectomycorrhizal septomycetes, including truffles. Microb Ecol 47:416–426<br />

Sinclair R (1990) Water relations in orchids. In: Arditti J (ed) Orchid biology, reviews and<br />

perspectives, vol V. Timber, Portland OR, pp 61–115<br />

Sturz AV, Christie BR, Nowak J (2000) Bacterial endophytes: potential role in developing<br />

sustainable systems of crop protection. Crit Rev Plant Prot 19:1–30<br />

Taylor DL, Bruns TD (1997) Independent, specialized invasions of ectomycorrhizal mutualism<br />

by two nonphotosynthetic orchids. Proc Natl Acad Sci USA 94:4510–4515<br />

Taylor DL, Bruns TD (1999) Population, habitat and genetic correlates of mycorrhizal<br />

specialization in the ‘cheating’ orchids Corallorhiza maculata and C. mertensiana.Mol<br />

Ecol 8:1719–1732<br />

Taylor DL, Bruns TD, Leake JR, Read DJ (2002) Mycorrhizal specificity and function<br />

in myco-heterotrophic plants. In: van der Heijden MGA, Sanders IR (eds) Mycorrhizal<br />

ecology Ecological studies vol. 157 Springer, Berlin Heidelberg New York,<br />

pp 375–414<br />

Taylor DL, Bruns TD, Szaro TM, Hodges SA (2003) Divergence in mycorrhizal specialization<br />

within Hexalectris spicata (Orchidaceae), a nonphotosynthetic desert orchid. Am J Bot<br />

90:1168–1179


9 Microbial Endophytes of Orchid Roots 177<br />

Terashita T (1985) Fungi inhabiting wild orchids in Japan (III). A symbiotic experiment<br />

with Armillariella mellea and Galeola septentrionalis. Trans Mycol Soc Jpn 26:47–53<br />

Tremblay RL, Zimmerman J, Lebrón LL, Bayman P, Sastre I, Axelrod F, Alers-García J<br />

(1998) Host specificity and low reproductive success in the rare endemic Puerto Rican<br />

orchidLepanthes caritensis. Biol Conserv 85:297–304<br />

Tsavkelova EA, Cherdyntseva TA, Lobakova ES, Kolomeitseva GL, Netrusov AI (2001) Microbiota<br />

of the orchid rhizoplane. Mikrobiologiya 70:567–573<br />

Tsavkelova EA, Lobakova ES, Kolomeitseva GL, Cherdyntseva TA, Netrusov AI (2003) Associative<br />

cyanobacteria isolated from the roots of epiphytic orchids. Mikrobiologiya<br />

72:105–10<br />

Umata H (1995) Seed germination of Galeola altissima, an achlorophyllous orchid, with<br />

aphyllophorales fungi. Mycoscience 36:369–372<br />

Umata H (1998) A new biological function of shiitake mushroom, Lentinula edodes, in<br />

a myco-heterotrophic orchid, Erythrorchis ochobiensis. Mycoscience 38:355–357<br />

Umata H (1997a) Formation of endomycorrhizas by an achlorophyllous orchid, Erythorchis<br />

ochobiensis,andAuricularia polytricha. Mycoscience 36:369–372<br />

Umata H (1997b) In vitro germination of Erythrorchis ochobiensis (Orchidaceae) in the<br />

presence of Lyophyllum shimeji, an ectomycorrhizal fungus. Mycoscience 38:335–339<br />

Umata H (1999) Germination and growth of Erythrorchis ochobiensis (Orchidaceae) accelerated<br />

by monokaryons and dikaryons of Lenzites betulinus and Trametes hirsuta.<br />

Mycoscience 40:367–371<br />

Vandenkoornhuyse P, Baldauf SL, Leyval C, Straczek J, Young JPW (2002) Extensive fungal<br />

diversity in plant roots. Science 295:2051<br />

Vujanovic V, St.-Arnaud M, Barabé D, Thibeault G (2000) Viability testing of orchid seed<br />

and the promotion of colouration and germination. Ann Bot 86:79–86<br />

Wilkinson KG, Dixon KW, Sivasithamparam K (1989) Interaction of soil bacteria, mycorrhizal<br />

fungi and orchid seed in relation to germination of Australian orchids. New<br />

Phytol 112:429–435<br />

Wilkinson KG, Sivasithamparam K, Dixon KW, Fahy PC, Bradley JK (1994) Identification<br />

and characterisation of bacteria associated with Western Australian orchids. Soil Biol<br />

Biochem 26:137–142<br />

Zelmer CD (1994) Interactions between northern terrestrial orchids and fungi in nature.<br />

MSc Thesis, University of Alberta, Edmonton, Alberta, Canada<br />

Zelmer CD, RS Currah (1995) Evidence of fungal liaison between Corallorhiza trifida (Orchidaceae)<br />

and Pinus contorta (Pinaceae). Can J Bot 73:862–866<br />

Zelmer CD, Cuthbertson L, Currah RS (1996) Fungi associated with terrestrial orchid<br />

mycorrhizas, seeds and protocorms. Mycoscience 37:439–448


10<br />

10.1<br />

Introduction<br />

Fungal Endophytes in Submerged Roots<br />

Felix Bärlocher<br />

It has long been known that plants harbour fungal endophytes, and it was<br />

suspected that systemic grass endophytes, primarily clavicipitaceous fungi,<br />

are associated with toxicity to grazing livestock (Saikkonen et al. 1998). This<br />

connection was firmly established in the 1970s (Bacon et al. 1977). The early<br />

emphasis on grasses and their endophytes have led some authors to consider<br />

the term endophyte as being synonymous with mutualist. However,<br />

many fungal pathogens may be latent in grasses without causing disease<br />

or long before the outbreak of disease symptoms (Petrini 1991; Fisher and<br />

Petrini 1993). The first systematic surveys of plants other than grasses<br />

were stimulated by the observation that many common phyllosphere fungi<br />

invade stomatal cavities of Douglas fir needles within their first year. Bernstein<br />

and Carroll (1977) demonstrated that with increasing age, all needle<br />

segments become infected with endophytes. The presence of primarily<br />

non-balansiaceous endophytes was extended to other conifers (Carroll et<br />

al. 1977) and has since been documented in every tree, shrub and herb<br />

that has been examined (Carroll 1995; Sridhar and Raviraja 1995; Saikkonnen<br />

et al. 1998). Generally, a large number of species can be isolated from<br />

a given host, yet only four to five are common and likely to be host specific<br />

(Fisher and Petrini 1993). Community ordination analyses have generally<br />

shown that endophyte assemblages are specific at the host species level,<br />

and may be impoverished outside the host’s natural range. While a few of<br />

these associations provide clear benefits to the plant by fungal interference<br />

with herbivores or microbial pathogens, others eventually cause damage<br />

to the plant, while some are essentially neutral. A widely accepted definition<br />

of an endophyte is as an agent of a currently asymptomatic infection,<br />

without specifying the role of the agent in the host or its development<br />

at a later stage (Petrini 1991; Fisher and Petrini 1993; Schulz et al. 1998).<br />

However, Schulz et al. (1999) showed that even in infections without visible<br />

symptoms, colonisation led to the synthesis of higher concentrations of<br />

Felix Bärlocher: 63B York Street, Department of Biology, Mount Allison University, Sackville,<br />

New Brunswick, E4L 1G7, Canada, E-mail: fbaerlocher@mta.ca<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


180 F. Bärlocher<br />

potentially antimicrobial compounds. In vitro, endophytic fungi produce<br />

more herbicidally active substances than soil fungi. Schulz et al. (1999)<br />

therefore hypothesise that the host-endophyte interaction is a case of balanced<br />

antagonism: pathogens overcome the host’s defences to the extent<br />

that they cause visible damage, whereas endophytic virulence is only sufficient<br />

to be able to infect and colonise without causing visible damage. If<br />

the balance shifts, the endophyte may turn pathogenic.<br />

Much of the current interest in endophytes is based on the hope of finding<br />

uniquesecondary metabolitesandenzymesaffectingplants,herbivoresand<br />

microbes, with potential applications in medicine and agriculture (Petrini<br />

et al. 1992). The continuum of endophyte interactions with plants also provides<br />

interesting case studies for the evolution of mutualism and pathology,<br />

and for understanding how environmental factors might favour one or the<br />

other (Carroll 1988). Until recently, it has commonly been believed that<br />

the first fungi were saprotrophs from which necrotrophs and biotrophs<br />

evolved, but there are convincing arguments supporting an alternative<br />

view (Parbery 1996). A recent comparison of 1,551 ribosomal sequences<br />

of the two sister groups of chitinous fungi, the Glomeromycota and the<br />

Dikaryomycota, both of which have symbiotic life-styles, suggests that the<br />

symbiosis between fungi and green plants was present before the colonisation<br />

of land by plants (Tehler et al. 2003). Finally, endophytes may not have<br />

been seriously taken into account when assessing diversity (Hawksworth<br />

1991, 2001); Dreyfuss and Chapela (1994) concluded that over 1.3 million<br />

species of fungal endophytes remain to be discovered and described.<br />

10.2<br />

Aquatic Hyphomycetes<br />

Up to 99% of the energy available to stream communities consists of<br />

terrestrial plant detritus (leaves, needles, twigs; Allan 1995). Aquatic hyphomycetes,<br />

a heterogeneous group of aquatic fungi, are an indispensable<br />

link in the food web between this detritus and stream invertebrates<br />

(Bärlocher 1992). The annual fungal production per stream bed area falls<br />

within the same order of magnitude as that of bacteria and invertebrates<br />

(Suberkropp 1997). Aquatic hyphomycetes disperse from leaf to leaf by producing<br />

conidia, whose shapes are predominantly tetraradiate (four arms)<br />

or sigmoid. Both types have been shown to increase the conidium’s likelihood<br />

of settling and germinating on new leaves; they are clearly the result<br />

of convergent evolution (Webster 1987).<br />

In temperate streams, the number of conidia in the water column declines<br />

from up to 30,000 l −1 in late fall to almost nil during summer,<br />

undoubtedly a response to the seasonal availability of terrestrial leaves


10 Fungal Endophytes in Submerged Roots 181<br />

(Bärlocher 1992, 2000). Combined with the unidirectional displacement<br />

of substrates and spores in running water, this raises the question of how<br />

aquatic hyphomycetes can maintain themselves within a given reach of<br />

a stream and avoid being washed downstream. Potential solutions include<br />

(1) the fact that the fungi also colonise woody substrates, which can persist<br />

for several years in a stream, (2) the presence of teleomorphs in some<br />

species with ascospores that may be dispersed aerially, (3) dispersal of<br />

fungal-colonised leaves or conidia by animals, (4) the occurrence of the<br />

fungi in terrestrial habitats, e.g. as plant pathogens or endophytes (Bärlocher<br />

1992). For example, Hartig (1880) first described a parasite of maple<br />

seedlings as Cercospora acerina. Later, its identity as Centrospora acerina<br />

was established, and it is now known as Mycocentrospora acerina (Hartig)<br />

Deighton. It is a remarkably widespread and versatile species: it is<br />

a well-known plant pathogen, has been implicated in human infections,<br />

andisacommonstreamfungus.Morphologically,thereisnodifference<br />

among various strains. Iqbal and Webster (1969) showed that strains they<br />

isolated from a stream were pathogenic to carrots and parsnips. Nemec<br />

(1969) isolated Anguillospora longissima (Sacc. & Syd.) Ingold and Tetracladium<br />

marchalianum de Wild., which had been isolated from aqueous<br />

habitats, from the roots of diseased strawberry plants. Several other species<br />

have also been isolated from terrestrial root surfaces of apparently healthy<br />

plants (Waid 1954; Taylor and Parkinson 1965; Parkinson and Thomas<br />

1969; Watanabe 1975). These observations suggested that some aquatic<br />

hyphomycetes might be root endophytes.<br />

10.3<br />

Fungi in Submerged Roots<br />

Fisher and Petrini (1989) were the first to demonstrate an endophytic phase<br />

of two aquatic hyphomycete species. They examined terrestrial roots of Alnus<br />

glutinosa (L.) Gaertner on the banks of Exeter Canal (Exeter, Devon,<br />

UK). Only 1.7 and 0.7% of 300 root segments were colonised by the aquatic<br />

hyphomycetes Tricladium splendens and Campylospora purvula, respectively,comparedtothe19%thatwerecolonisedbythemostcommon<br />

endophyte Cylindrocarpon destructans In a later study, Fisher et al. (1991)<br />

compared aquatic and terrestrial alder roots along the banks of the River<br />

Dart (Devon, UK) They separated roots into bark and xylem (decorticated<br />

roots), and found more endophytic aquatic hyphomycetes in the former.<br />

Mean frequency of occurrence of aquatic species in submerged roots was<br />

as high as 30%, compared to 12% on terrestrial roots. In addition to typical<br />

aquatic hyphomycetes, they also found species of the genera Fusarium<br />

and Cylindrocarpon. Members of these two taxa are often found on leaves


182 F. Bärlocher<br />

in decaying streams. Cluster and correspondence analyses suggested that<br />

aquatic and soil root samples are colonised by two distinct endophyte<br />

populations, indicating that the external environment may have a greater<br />

influence on endophyte communities of roots than those of leaves (Fisher<br />

and Petrini 1993). Three previously unknown species (Fontanospora fusiramos,<br />

and two species belonging to Filosporella) were subsequently isolated<br />

and described from submerged alder roots (Marvanová and Fisher 1991;<br />

Marvanová et al. 1992, 1997).<br />

The host range of endophytic aquatic hyphomycetes was extended by<br />

Sridhar and Bärlocher (1992a). They found additional species in spruce<br />

(Picea glauca [Moench] Voss), birch (Betula papyrifera Marsh) and maple<br />

(Acer spicatum Lam.). Again, fungal endophytes were more common in the<br />

bark,suggestingthatrootsarecolonisedbyfungisettlingonsurfacesand<br />

growing toward the interior. In addition to plating out surface-sterilised<br />

root fragments, Sridhar and Bärlocher (1992a) aerated them in distilled<br />

water and were able to observe release of typical tetraradiate or sigmoid<br />

conidia. However, aeration had to continue for 4 days (compared to the<br />

usual 1–2 days) before spores were detected (any superficial mycelia that<br />

may have been present were killed by surface sterilisation). Spore production<br />

per unit mass was less than 1 mg −1 , compared to 100–150 mg −1 from<br />

dead submerged branches, and up to 8,000 mg −1 on dead leaves (Gessner<br />

et al. 2003). Nevertheless, the root biomass in streams is considerable and<br />

its turnover rapid (Waid 1974), suggesting that it may be an important secondary<br />

resource for aquatic hyphomycetes. Their existence as endophytes<br />

may provide them with a head start in the use of root detritus, a possible<br />

advantage of the endophytic life style that has also been suggested for<br />

leaf-decomposing saprobes (Fisher and Petrini 1993).<br />

On spruce roots, the number aquatic hyphomycetes in the xylem was<br />

highest in 4- to 5-year-old segments (Sridhar and Bärlocher 1992b). Total<br />

fungal biomass, estimated by ergosterol, amounted to 0.002 to 0.2% of<br />

root biomass. This compares to values exceeding 15% on decaying leaves<br />

(Gessner et al. 2003).<br />

Iqbal et al. (1994) reported 17 species of endophytic aquatic hyphomycetes<br />

from tree roots along canal banks in Pakistan (Mangifera indica L.,<br />

Populus hybrida Reichb., Salix babylonica L.). Two plantation crops (Coffea<br />

arabica Linn., Hevea brasiliensis M.) and four ferns (Diplazium esculentum<br />

(Retz) Sw., Macrothelypteris torresiana (Gaudich.) Ching., Angiopteris<br />

evecta (Forst) Hoffin., Christela dentata Brownsey & Jermy), all from India,<br />

were also shown to harbour aquatic endophytes in their submerged roots<br />

(Raviraja et al. 1996). Again, their incidence was higher in the bark than in<br />

the xylem of tree roots. Conidium release per root biomass upon aeration<br />

was much higher from the fern C. dentata (12,900 g −1 ) than from the other<br />

plants (36–410 g −1 ).


10 Fungal Endophytes in Submerged Roots 183<br />

Permanently or periodically submerged roots are common in mangrove<br />

swamps. Ananda and Sridhar (2002) examined fungal epiphytes in prop<br />

rootsorpneumatophores(whichcyclebetweenexposuretoairandimmersion<br />

in salt or brackish water) of Avicennia officinalis L., Rhizophora mucronata<br />

Lamk. and Sonneratia caseolaris (L.) Engl.. Aquatic hyphomycetes<br />

were represented by Mycocentrospora acerina in roots of A. officinalis L.,<br />

and Triscelophorus acuminatus in roots of R. mucronata and S. caseolaris.<br />

In addition, various marine and terrestrial fungi were found.<br />

Overall, 35 aquatic hyphomycete species (plus seven taxa identified to<br />

genus) have been reported from submerged roots of 13 plants, including<br />

Angiosperms, Gymnosperms and ferns in eight studies (Table 10.1).<br />

This corresponds to roughly 10% of the total number of described species<br />

(L. Marvanová, personal communication). Clearly, submerged roots can<br />

provide a stationary refuge for aquatic hyphomycetes, which may help<br />

them maintain their presence in a given stream reach despite the unidirectional<br />

flow of water.<br />

Table 10.1. Endophytic aquatic hyphomycetes recovered from roots, submerged in saltwater<br />

(*) or freshwater (all others). Root sections: R Entire root, B bark, X xylem (decorticated<br />

root)<br />

Fungus Substrate<br />

Anguillospora filiformis<br />

Greath.<br />

A. longissima<br />

(de Wild.) Ingold<br />

Articulospora antipodea<br />

Roldán<br />

A. atra<br />

Descals<br />

A. tetracladia<br />

Ingold<br />

A. proliferata<br />

Jooste, Radon & Merwe<br />

Bacillispora inflata<br />

Iqbal & Bhatty<br />

Root<br />

section <strong>References</strong><br />

Acer spicatum B, X Raviraja et al. 1996;<br />

Sridhar and Bärlocher 1992b<br />

Betula papyrifera B, X Sridhar and Bärlocher 1992a<br />

Picea glauca B Sridhar and Bärlocher 1992a<br />

Mangifera indica B, X Iqbal et al. 1995<br />

Populus hybrida B, X Iqbal et al. 1995<br />

Salix babylonica B, X Iqbal et al. 1995<br />

Picea glauca B, X Sridhar and Bärlocher 1992a<br />

Alnus glutinosa B Fisher et al. 1991<br />

Picea glauca B, X Sridhar and Bärlocher 1992a<br />

Acer spicatum B, X Fisher et al. 1991<br />

Alnus glutinosa B, X Fisher et al. 1991; Sridhar<br />

and Bärlocher 1992a, 1992b<br />

Picea glauca B, X Sridhar and Bärlocher 1992a<br />

Mangifera indica B, X Iqbal et al. 1995<br />

Populus hybrida B, X Iqbal et al. 1995<br />

Salix babylonica B, X Iqbal et al. 1995<br />

Mangifera indica B Iqbal et al. 1995<br />

Populus hybrida B Iqbal et al. 1995<br />

Salix babylonica B Iqbal et al. 1995


184 F. Bärlocher<br />

Table 10.1. (continued)<br />

Fungus Substrate<br />

Campylospora<br />

chaetocladia<br />

Ranzoni<br />

Clavariopsis aquatica<br />

de Wild.<br />

Root<br />

section <strong>References</strong><br />

Salix babylonica B Iqbal et al. 1995<br />

Alnus glutinosa B, X Fisher et al. 1991<br />

Picea glauca X Sridhar and Bärlocher 1992a<br />

Populus hybrida B Iqbal et al. 1995<br />

Salix babylonica B Iqbal et al. 1995<br />

C. azlanii Nawawi Mangifera indica B Iqbal et al. 1995<br />

Cylindrocarpon aquaticum<br />

(Nils.) Marvanová &<br />

Descals<br />

Acer spicatum B, X Sridhar and Bärlocher 1992a<br />

Mangifera indica B, X Iqbal et al. 1995<br />

Picea glauca B Sridhar and Bärlocher 1992a<br />

Populus hybrida B, X Iqbal et al. 1995<br />

Salix babylonica B, X Iqbal et al. 1995<br />

Filosporella sp. Alnus glutinosa B Fisher et al. 1991<br />

F. fistucella<br />

Marvanová & Fisher<br />

Alnus glutinosa B Marvanová and Fisher 1991<br />

F. versimorpha<br />

Marvanová et al.<br />

Alnus glutinosa B Marvanová et al. 1992<br />

Flagellospora curvula<br />

Ingold<br />

F. fusarioides<br />

Iqbal<br />

F. penicillioides<br />

Ingold<br />

Mangifera indica B Iqbal et al. 1995<br />

Populus hybrida B Iqbal et al. 1995<br />

Salix babylonica B, X Iqbal et al. 1995<br />

Mangifera indica B, X Iqbal et al. 1995<br />

Populus hybrida B, X Iqbal et al. 1995<br />

Salix babylonica B, X Iqbal et al. 1995<br />

Mangifera indica B, X Iqbal et al. 1995<br />

Populus hybrida B, X Iqbal et al. 1995<br />

Salix babylonica B, X Iqbal et al. 1995<br />

Fontanospora fusiramosa<br />

Marvanova et al.<br />

Alnus glutinosa R Marvanová et al. 1997<br />

Geniculospora sp. Picea glauca B, X Sridhar and Bärlocher 1992b<br />

Heliscus lugdunensis<br />

Sacc. & Therry<br />

Lunulospora curvula<br />

Ingold<br />

Acer spicatum B, X Sridhar and Bärlocher 1992a<br />

Alnus glutinosa B, X Iqbal et al. 1995<br />

Betula papyrifera B, X Sridhar and Bärlocher 1992a<br />

Picea glauca B, X Sridhar and Bärlocher<br />

1992a, 1992b<br />

Salix babylonica B,X Iqbal et al. 1995<br />

Alnus glutinosa B Fisher et al. 1991<br />

Angiopteris evecta R Raviraja et al. 1996<br />

Christela dentata R Raviraja et al. 1996<br />

Coffee arabica B, X Raviraja et al. 1996


10 Fungal Endophytes in Submerged Roots 185<br />

Table 10.1. (continued)<br />

Fungus Substrate<br />

Root<br />

section <strong>References</strong><br />

Hevea brasiliensis X Raviraja et al. 1996<br />

Mangifera indica B Iqbal et al. 1995<br />

Populus hybrida B Iqbal et al. 1995<br />

Salix babylonica B Iqbal et al. 1995<br />

Mycocentrospora sp. 1 Alnus glutinosa B Fisher et al. 1991<br />

Mycocentrospora sp. 2 Acer spicatum B, X Sridhar and Bärlocher 1992a<br />

Picea glauca B, X Sridhar and Bärlocher 1992a<br />

Mycocentrospora sp. 3 Coffee arabica B Raviraja et al. 1996<br />

Diplazium esculentum R Raviraja et al. 1996<br />

Hevea brasiliensis X Raviraja et al. 1996<br />

Macrothelypteris<br />

torresiana<br />

R Raviraja et al. 1996<br />

M. acerina<br />

(Hartig) Deighton<br />

*Avicennia officinalis R Ananda and Sridhar 2002<br />

M. clavata Iqbal Betula papyrifera B, X Sridhar and Bärlocher 1992a<br />

Picea glauca B, X Sridhar and Bärlocher 1992a<br />

M. iqbalii sp.ind.F.BareenMangifera indica B, X Iqbal et al. 1995<br />

Salix babylonica B, X Iqbal et al. 1995<br />

Phalangispora constricta<br />

Nawawi & Webster<br />

Picea glauca B, X Sridhar and Bärlocher 1992b<br />

Pseudoanguillospora sp. Alnus glutinosa B Fisher et al. 1991<br />

Tetrabrachium elegans Acer spicatum B, x Sridhar and Bärlocher 1992a<br />

Nawawi & Kuthubutheen Betula papyrifera B, X Sridhar and Bärlocher 1992a<br />

Picea glauca B Sridhar and Bärlocher 1992a<br />

Tetracladium sp Angiopteris evecta R Raviraja et al. 1996<br />

T. furcatum Descals Angiopteris evecta R Raviraja et al. 1996<br />

T. marchalianum Mangifera indica B Iqbal et al. 1995<br />

de Wild.<br />

Populus hybrida B Iqbal et al. 1995<br />

Salix babylonica B Iqbal et al. 1995<br />

T. setigerum<br />

(Grove) Ingold<br />

Picea glauca B Sridhar and Bärlocher 1992b<br />

Tricladium chaetocladium Alnus glutinosa B Fisher et al. 1991<br />

Ingold<br />

Alnus glutinosa B Fisher et al. 1991<br />

Tricellula aquatica<br />

Webster<br />

Mangifera indica B Iqbal et al. 1995<br />

Triscelophorus acuminatus Angiopteris evecta R Raviraja et al. 1996<br />

Nawawi<br />

Christela dentata R Raviraja et al. 1996<br />

Coffea arabica B Raviraja et al. 1996<br />

Diplazium esculatum R Raviraja et al. 1996<br />

Hevea brasiliensis B, X Raviraja et al. 1996


186 F. Bärlocher<br />

Table 10.1. (continued)<br />

Fungus Substrate<br />

T. konajensis<br />

Sridhar & Kaveriappa<br />

T. monosporus<br />

Ingold<br />

Tumularia aquatica<br />

(Ingold)<br />

Marvanová & Descals<br />

Varicosporium elodeae<br />

Kegel<br />

Root<br />

section <strong>References</strong><br />

Macrothelypteris<br />

torresiana<br />

R Raviraja et al. 1996<br />

*Rhizophora<br />

mucronata<br />

R Ananda and Sridhar 2002<br />

*Sonneratia caseolaris R Ananda and Sridhar 2002<br />

Antipteris evecta R Raviraja et al. 1996<br />

Christela dentata R Raviraja et al. 1996<br />

Coffea arabica B Raviraja et al. 1996<br />

Macrothylpteris<br />

torresiana<br />

R Raviraja et al. 1996<br />

Angiopteris evecta R Raviraja et al. 1996<br />

Christela dentata R Raviraja et al. 1996<br />

Coffea arabica B, X Raviraja et al. 1996<br />

Diplazium esculatum R Raviraja et al. 1996<br />

Macrothelypteris R Raviraja et al. 1996<br />

torresiana<br />

Mangifera indica B Iqbal et al. 1995<br />

Populus hybrida B Iqbal et al. 1995<br />

Salix babylonica B Iqbal et al. 1995<br />

Alnus glutinosa B Fisher et al. 1991<br />

Alnus glutinosa B Fisher et al. 1991<br />

Picea glauca B, X Sridhar and Bärlocher<br />

1992a, 1992b<br />

V. giganteum Crane Picea glauca B, X Sridhar and Bärlocher<br />

1992a, 1992b<br />

10.4<br />

Conclusions and Outlook<br />

Work on submerged roots, primarily in fresh water, has been dominated<br />

by a very specific objective: to evaluate their role as habitat for aquatic<br />

hyphomycetes. Other aspects of the plant-endophyte relationship include<br />

the potential production of unique secondary metabolites allowing the<br />

fungi to live within the plant without overt symptoms, and which might<br />

be toxic to potential pathogens or herbivores. Several observations suggest<br />

that aquatic fungi can produce diffusible antibiotics. For example,<br />

Massarina aquatica, the teleomorph of Tumularia aquatica, releases antifungal<br />

substances (Fisher and Anson 1983). Similar observations on other


10 Fungal Endophytes in Submerged Roots 187<br />

species have been reported by Asthana and Shearer (1990) and Poch et<br />

al. (1992). Chamier et al. (1984) demonstrated inhibition of bacteria by<br />

aquatic hyphomycetes in field experiments. Isolation and characterisation<br />

of antimicrobial compounds from Anguillospora longissima and A. crassa<br />

resulted in the discovery of novel metabolites (Harrigan et al. 1995). Two<br />

surveys of aquatic hyphomycetes and ascomycetes demonstrated that antibacterial<br />

and antifungal substances are produced by about one-half of<br />

the species tested (Gulis and Stephanovich 1999; Shearer and Zare-Maivan<br />

1988). Lignicolous aquatic ascomycetes and hyphomycetes were generally<br />

more antagonistic than foliicolous species, possibly because long-lasting<br />

substrata,suchaswood,favourcolonisationbyspeciescapableofdefending<br />

captured resources (Shearer 1992). It is currently unknown how such compounds<br />

affect the root’s susceptibility toward pathogens or herbivores.<br />

Sexual and asexual reproduction of the endophyte are often initiated<br />

upon the death of the host tissue (Fisher et al. 1986). Sridhar and Bärlocher<br />

(1992a) reported that Heliscus lugdunensis produced a teleomorph upon<br />

subculturing; aquatic hyphomycetes endophytic in roots may therefore<br />

be useful in establishing additional anamorph-teleomorph connections<br />

(Webster 1992; Sivichai and Jones 2003).<br />

It is generally accepted that aquatic hyphomycetes and ascomycetes had<br />

terrestrial ancestors, and several have indeed close terrestrial relatives<br />

(Kong et al. 2000; Liew et al. 2002). Shearer (1993) suggested that when<br />

terrestrial plants invaded freshwater habitats, they brought with them fungalpathogens,endophytesandsaprobes.Alternatively,plantdetritus,precolonised<br />

by fungal biotrophs or saprotrophs may have fallen into streams.<br />

Some of these fungi may subsequently have adapted to dispersal and reproduction<br />

in water. The most comprehensive analysis of fungal gene sequences<br />

suggests that the biotrophic lifestyle was a synapomorphic trait<br />

(i.e. present in a common ancestor; Tehler et al. 2003). The first fungi that<br />

colonised the terrestrial habitat most likely did so while closely associated<br />

with plants. The question remains whether such fungi first evolved into<br />

terrestrial saprobes, and then into aquatic hyphomycetes, or whether there<br />

was a direct transition from terrestrial biotrophs to aquatic saprobes. Were<br />

the presumed ancestors restricted to specific plant organs, e.g. aerial twigs<br />

and leaves, or roots? Upon their death, roots submerged in streams may<br />

have released propagules of fungal endophytes, some of which settled on<br />

other types of imported terrestrial detritus, such as leaves. Over time, this<br />

may have favoured adaptation to life in running water. Or, terrestrial leaves<br />

infected with endophytes were shed and landed in streams. Even waterfilms<br />

on soil or between layers of terrestrial leaf layer may have selected<br />

for tetraradiate spore shapes (Bandoni 1975) and predisposed some fungi<br />

for their eventual evolution into aquatic hyphomycetes. There are several<br />

reports of aquatic hyphomycete conidia in rainwater dripping from trees


188 F. Bärlocher<br />

(Ando and Tubaki 1984; Bärlocher 1992; Czeczuga and Orlowska 1999), and<br />

Widler and Müller (1984) isolated two undescribed species of Gyoerffyella<br />

and Varicosporium from green twigs. Iqbal et al. (1995) reported 14 species<br />

from submerged green leaves.<br />

It will be of considerable interest to investigate the common route of<br />

invasion by aquatic hyphomycetes: do they first colonise the roots, and<br />

then spread through the rest of the plant, or do some invade aerial parts?<br />

A thorough study of endophytes in aerial and subterranean plant parts at<br />

various ages, and molecular data of such strains may eventually allow us to<br />

reconstruct the origins of aquatic hyphomycetes.<br />

<strong>References</strong><br />

Allan JD (1995) Stream ecology. Chapman and Hall, London<br />

Ananda K, Sridhar KR (2002) Diversity of endophytic fungi in the roots of mangrove species<br />

on the west coast of India. Can J Microbiol 48:871–878<br />

Ando K, Tubaki K (1984) Some undescribed hyphomycetes in the rain drops from intact<br />

leaf-surface. Trans Mycol Soc Jpn 25:21–37<br />

Asthana A, Shearer CA (1990) Antagonistic activity of Pseudohalonectria and Ophioceras.<br />

Mycologia 82:554–561<br />

Bacon CW, Porter JK, Robbins JD, Luttrell ES (1977) Epichloë typhina from toxic tall fescue<br />

grasses. Appl Environ Microbiol 34:576–581<br />

Bandoni RJ (1975) Significance of the tetraradiate form in dispersal of terrestrial fungi. Rep<br />

Tottori Mycol Inst 12:105–113<br />

Bärlocher F (1992) Research on aquatic hyphomycetes: historical background and overview.<br />

In: Bärlocher F (ed) The ecology of aquatic hyphomycetes Springer, Berlin Heidelberg<br />

New York, pp 1–15<br />

Bärlocher F (2000) Water-borne conidia of aquatic hyphomycetes: seasonal and yearly<br />

patterns in Catamaran Brook, New Brunswick, Canada. Can J Bot 78:157–167<br />

Bernstein ME, Carroll G (1977) Internal fungi in old-growth Douglas fir foliage. Can J Bot<br />

55:644–653<br />

Carroll GC (1988) Fungal endophytes in stems and leaves: from latent pathogen to mutualistic<br />

symbiont. Ecology 69:2–9<br />

Carroll GC (1995) Forest endophytes: pattern and process. Can J Bot 73 [Suppl 1]:S1316–<br />

S1324<br />

Carroll FE, Müller E, Sutton BC (1977) Preliminary studies on the incidence of needle<br />

endophytes in some European conifers. Sydowia 29:87–103<br />

Chamier A-C, Dixon P, Archer SA (1984) The spatial distribution of fungi on decomposing<br />

alder leaves in a freshwater stream. Oecologia 64:92–103<br />

Czeczuga B, Orlowska M (1999) Hyphomycetes in rain water, melting snow and ice. Acta<br />

Mycol 34:181–200<br />

Dreyfuss MM, Chapela ICH (1994) Potential of fungi in the discovery of novel, low-molecular<br />

weight pharmaceuticals. In Gullo VP (ed) The discovery of natural products with therapeutic<br />

potential. Butterworth-Heinemann, Boston, pp 49–80<br />

Fisher PJ, Anson AE (1983) Antifungal effects of Massarina aquatica growing on oak wood.<br />

Trans Br Mycol Soc 81:523–527<br />

Fisher PJ, Petrini O (1989) Two aquatic hyphomycetes as endophytes in Alnus glutinosa<br />

roots. Mycol Res 92:367–368


10 Fungal Endophytes in Submerged Roots 189<br />

Fisher PJ, Petrini O (1993) Ecology, biodiversity and physiology of endophytic fungi. Curr<br />

Top Bot Res 1:271–279<br />

Fisher PJ, Anson AE, Petrini O (1986) Fungal endophytes in Ulex europaeus and Ulex gallii.<br />

Trans Brit Mycol Soc 86:153–156<br />

Fisher PJ, Petrini O, Webster J (1991) Aquatic hyphomycetes and other fungi in living aquatic<br />

and terrestrial roots of Alnus glutinosa. Mycol Res 95:543–547<br />

Gessner MO, Bärlocher F, Chauvet E (2003) Qualitative and quantitative analyses of aquatic<br />

hyphomycetes in streams. Fungal Div Res Ser 10:127–157<br />

Gulis VI, Stephanovich AI (1999) Antibiotic effects of some aquatic hyphomycetes. Mycol<br />

Res 103:111–115<br />

Harrigan GG, Armentrout GL, Gloer JB, Shearer CA (1995) Anguillosporal, a new antibacterial<br />

and antifungal metabolite from the freshwater fungus Anguillospora longissima.<br />

J Nat Prod 58:1467–1469<br />

Hartig R (1880) Der Ahornkeimlingspilz, Cercospora acerina. Untersuch Forstbot Inst<br />

München 1:58–61<br />

Hawksworth DL (1991) The fungal dimension of biodiversity: magnitude, significance and<br />

conservation. Mycol Res 95:641–655<br />

Hawksworth DL (2001) The magnitude of fungal diversity: the 1.5 million species estimate<br />

revisited. Mycol Res 105:1422–1431<br />

Iqbal SH, Webster J (1969) Pathogenicity of aquatic isolates of Centrospora acerina to carrots<br />

and parsnips. Trans Br Mycol Soc 53:486–490<br />

Iqbal SH, Firdaus-e-Bareen, Yousaf N (1994) Freshwater hyphomycete communities in<br />

acanal.1.Endophytichyphomycetesofsubmergedrootsoftreesshelteringacanal<br />

bank. Can J Bot 73:538–543<br />

Iqbal SH, Akhtar G, Firdaus-e-Bareen (1995) Endophytic freshwater hyphomycetes of submerged<br />

leaves of some plants lining a canal bank. Pak J Plant Sci 1:239–254<br />

Kong RY, Chan JY, Mitchell JI, Vrijmoed LLP, Jones EBG (2000) Relationships of<br />

Halosarpheia, Lignincola and Nais inferred from partial 18S rDNA. Mycol Res 103:1399–<br />

1403<br />

Liew ECY, Aptroot A, Hyde KD (2002) An evaluation of the monophyly of Massarina based<br />

on ribosomal DNA sequences. Mycologia 94:803–813<br />

Marvanová L, Fisher PJ (1991) A new endophytic hyphomycete from alder roots. Nova<br />

Hedwigia 52:33–37<br />

Marvanová L, Fisher PJ, Aimer R, Segedin BC (1992) A new Filosporella from alder roots<br />

and from water. Nova Hedwigia 54:151–158<br />

Marvanová L, Fisher PJ, Descals E, Bärlocher F (1997) Fontanospora fusiramosa sp. nov.,<br />

a hyphomycete from live tree roots and from stream foam. Czech Mycol 50:3–11<br />

Nemec S (1969) Sporulation and identification of fungi isolated from root-rot in diseased<br />

strawberry plants. Phytopathology 59:1552–1553<br />

Parbery DG (1996) Trophism and the ecology of fungi associated with plants. Biol Rev<br />

71:473–527<br />

Parkinson D, Thomas S (1969) Studies on fungi in the root region. Plant Soil 31:299–310<br />

Petrini O (1991) Fungal endophytes of tree leaves. In: Andrews J, Hirano S (eds) Microbial<br />

ecology of leaves. Springer, Berlin Heidelberg New York, pp 179–197<br />

Petrini O, Sieber TH, Toti L, Viret O (1992) Ecology, metabolite production, and substrate<br />

utilization in endophytic fungi. Nat Toxins 1:185–196<br />

Poch GK, Gloer JB, Shearer CA (1992) New bioactive metabolites from a freshwater isolate<br />

of the fungus Kirschsteiniothelia sp. J Nat Prod 55:1093–1099<br />

Raviraja NS, Sridhar KR, Bärlocher F (1996) Endophytic aquatic hyphomycetes of roots<br />

from plantation crops and ferns from India. Sydowia 48:152–160


190 F. Bärlocher<br />

Saikkonen K, Faeth SH, Helander M, Sullivan TJ (1998) Fungal endophytes: a continuum of<br />

interactions with host plants. Annu Rev Ecol Syst 29:319–343<br />

Schulz B, Guske S, Dammann U, Boyle C (1998) Endophyte-host interactions. II. Defining<br />

symbiosis of the endophyte-host interaction. Symbiosis 25:213–227<br />

Schulz B, Römmert A-K, Damman U, Aust H-J, Strack D (1999) The endophyte-host interaction:<br />

a balanced antagonism? Mycol Res 103:1275–1283<br />

Shearer CA (1992) The role of woody debris. In: Bärlocher F (ed) The ecology of aquatic<br />

hyphomycetes. Springer, Berlin Heidelberg New York, pp 77–98<br />

Shearer CA (1993) The freshwater ascomycetes. Nova Hedwigia 56:1–33<br />

Shearer CA, Zare-Maivan H (1988) In vitro hyphal interactions among wood- and leafinhabiting<br />

ascomycetes and fungi imperfecti from freshwater habitats. Mycologia<br />

80:31–37<br />

Sivichai S, Jones EBG (2003) Teleomorphic-anamorphic connections of freshwater fungi.<br />

Fung Div Res Ser 10:259–272<br />

Sridhar KR, Bärlocher F (1992a) Endophytic aquatic hyphomycetes of roots from spruce,<br />

birch and maple. Mycol Res 96:305–308<br />

Sridhar KR, Bärlocher F (1992b) Aquatic hyphomycetes in spruce roots. Mycologia<br />

84:580–584<br />

Sridhar KR, Raviraja NS (1995) Endophytes – a crucial issue. Curr Sci 69:570–571<br />

Suberkropp K (1997) Annual production of leaf-decaying fungi in a woodland stream.<br />

Freshwater Biol 38:169–178<br />

Taylor GS, Parkinson D (1965) Studies in the root region. IV. Fungi associated with the roots<br />

of Phaseolus vulgaris L. Plant Soil 22:1–20<br />

Tehler A, Little DP, Farris JS (2003) The full-length phylogenetic tree from 1551 ribosomal<br />

sequences of chitinous fungi. Mycol Res 107:901–916<br />

Waid JS (1954) Occurrence of aquatic hyphomycetes upon the root surfaces of beech grown<br />

in woodland soils. Trans Br Mycol Soc 37:420–421<br />

Waid JS (1974) Decomposition of roots. In: Dickinson CH, Pugh GJF (eds) Biology of plant<br />

litter decomposition. Academic, New York, pp 175–211<br />

Watanabe T (1975) Tetracladiumsetigerum, an aquatic hyphomycete associated with gentian<br />

and strawberry roots. Trans Mycol Soc Jpn 16:348–350<br />

Webster J (1987) Convergent evolution and the functional significance of spore shape in<br />

aquatic and semi-aquatic fungi. In: Rayner ADM, Brasier CM, Moore D (eds) Evolutionary<br />

biology of the fungi. Cambridge University Press, Cambridge, pp 191–201<br />

Webster J (1992) Anamorph-teleomorph relationships. In: Bärlocher F (ed) The ecology of<br />

aquatic hyphomycetes. Springer, Berlin Heidelberg New York, pp 99–117<br />

Widler B, Müller E (1984) Untersuchungen über endophytische Pilze von Arctostaphylos<br />

uva-ursi (L.) Sprengel (Ericaceae). Bot Helv 94:307–337


11<br />

11.1<br />

Introduction<br />

Nematophagous Fungi<br />

as Root Endophytes<br />

Luis V. Lopez-Llorca, Hans-Börje Jansson,<br />

José Gaspar Maciá Vicente, Jesús Salinas<br />

Nematophagous fungi constitute a group of fungal antagonists to nematodes.<br />

The latter are small roundworms living in soil and water. Most<br />

nematodes are saprotrophic, but many species are parasites of plants and<br />

animals (Poinar 1983). The nematophagous fungi have been suggested as<br />

promising candidates for biological control of parasitic nematodes (Stirling<br />

1991), but so far no successful commercial products have been presented.<br />

Many of the previous studies on these organisms have been concerned<br />

with the ecology and physiology of interactions between nematophagous<br />

fungi and nematodes. More recently, molecular techniques have been employed<br />

(Jansson and Lopez-Llorca 2001). Nematophagous fungi also have<br />

the ability to infect and colonise other organisms, including other fungi<br />

and plant roots (Jansson and Lopez-Llorca 2004). In the current review we<br />

will briefly describe the nematophagous fungi, with special emphasis on<br />

their interactions with plant roots.<br />

11.2<br />

Nematophagous Fungi<br />

Nematophagous fungi, or nematode-destroying fungi, have the capacity<br />

to infect, kill and digest living stages of their nematode hosts (eggs, juveniles<br />

and adults). These fungi are ubiquitous soil inhabitants found in<br />

most parts of the world and in all climate types (Barron 1977). Many of the<br />

Luis V. Lopez-Llorca: Departamento de Ciencias del Mar y Biología Aplicada, Universidad<br />

de Alicante, Apdo. 99, 03080 Alicante, Spain, E-mail: lv.lopez@ua.es<br />

Hans-Börje Jansson: Departamento de Ciencias del Mar y Biología Aplicada, Universidad<br />

de Alicante, Apdo. 99, 03080 Alicante, Spain, E-mail: hb.jansson@ua.es<br />

José Gaspar Maciá Vicente: Departamento de Ciencias del Mar y Biología Aplicada, Universidad<br />

de Alicante, Apdo. 99, 03080 Alicante, Spain, E-mail: jgmv@ua.es<br />

Jesús Salinas: Departamento de Ciencias del Mar y Biología Aplicada, Universidad de Alicante,<br />

Apdo. 99, 03080 Alicante, Spain, E-mail: jesus.salinas@ua.es<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


192 L.V. Lopez-Llorca et al.<br />

nematophagous fungi are facultative parasites and can grow saprophytically<br />

in soil. In the presence of hosts they can change from a saprophytic to<br />

a parasitic stage and form infection structures, e.g. trapping organs, hyphal<br />

coils or appressoria. These infection structures vary depending on the type<br />

of host–nematode, fungus or plant.<br />

Entomopathogenic fungi, e.g. Lecanicillium lecanii,havethecapacityto<br />

infect both nematode eggs (Meyer 1998) and other fungi. Furthermore,<br />

species closely related to nematophagous fungi, e.g. Arthrobotrys ferox,<br />

can infect other small soil animals like springtails (Rubner 1996), but are<br />

generally not known to infect nematodes.<br />

11.2.1<br />

Nematode Parasites<br />

The nematophagous fungi can be divided into four groups depending on<br />

their mode of attacking their hosts. The first three groups infect vermiform<br />

nematodes (juveniles and adults), whereas the fourth group infects<br />

nematode females and eggs (Jansson and Lopez-Llorca 2001). Nematodetrapping<br />

fungi use various types of trapping organs formed on their hyphae,<br />

e.g. adhesive networks, adhesive knobs (Fig. 11.1a) or constricting rings,<br />

and these fungi are facultative parasites to various extents. The nematodes<br />

are captured in the traps formed by the fungi either by adhesion<br />

or mechanical function. In the endoparasitic fungi, the spores (conidia,<br />

zoospores) function as infection structures, which either adhere to the nematode<br />

cuticle or are ingested. These fungi are generally obligate parasites<br />

of nematodes (Fig. 11.1b). The toxin-producing fungi, comprising for instance<br />

the common wood-decomposing oyster mushroom, intoxicate their<br />

nematode victims before penetrating them. The egg- and female parasitic<br />

attack mature females of cyst- and root-knot nematodes and the eggs they<br />

contain (Fig. 11.1c). Infection usually takes place via appressoria. Common<br />

to all types of nematophagous fungi is that after contact with the nematode<br />

cuticle, or egg shell, penetration takes place followed by digestion of the<br />

contents resulting in formation of new fungal biomass inside, and later<br />

outside, the nematode.<br />

Taxonomy<br />

The nematophagous fungi are found in most fungal taxa (Dackman et<br />

al. 1992). In the Basidiomycetes, nematophagous fungi such as the oyster<br />

mushroom (Pleurotus ostreatus) andHohenbuehelia spp. (teleomorph of<br />

Nematoctonus spp.) can be found. Many of the nematode-trapping fungi<br />

belong to the Deuteromycetes or mitosporic fungi, e.g. Arthrobotrys spp.<br />

and Monacrosporium spp., but the Arthrobotrys spp. have been found


11 Nematophagous Fungi as Root Endophytes 193<br />

Fig.11.1. a Adhesive knob trap of Monacrosporium haptotylum adhering to the nematode<br />

cuticle. Note adhesive pad between trap and nematode (arrow). b Conidia of the endoparasitic<br />

fungus Drechmeria coniospora adhering to the mouth region of a nematode. (From<br />

Jansson and Nordbring-Hertz 1983, courtesy of Society of General Microbiology). c A nematode<br />

egg infected by the egg-parasitic fungus Pochonia rubescens. (From Lopez-Llorca<br />

and Claugher 1990, courtesy of Elsevier). Bars a, b 5µm;c 4µm<br />

to be Ascomycetes with their teleomorph in Orbilia spp. (Pfister 1997).<br />

Other nematode-trapping fungi, e.g. Stylopage and Cystopage spp. are Zygomycetes,<br />

and some endoparasitic fungi, such as Catenaria anguillulae,are<br />

zoosporic Chytridiomycetes. The main egg-parasitic fungi are now placed<br />

in the new genus Pochonia (formerly Verticillium) (Gams and Zare 2003).<br />

Therefore, the various nematophagous fungi seem to have acquired their<br />

nematophagous ability independently during the course of evolution.<br />

The few phylogenetic studies that have been presented (Ahrén et al.<br />

1998; Hagedorn and Scholler 1999) show that the orbiliaceous nematodetrapping<br />

fungi are closely related, and that the type of trap, rather than<br />

traditional spore morphology, is more determinate on the species level.<br />

Biology<br />

We will focus on two types of nematophagous fungi: the nematode-trapping<br />

Arthrobotrys oligospora and the egg parasite Pochonia chlamydosporia.<br />

These fungi are common soil inhabitants living both saprophytically and<br />

parasitically and, as we will show, also endophytically.<br />

Arthrobotrys oligosporaforms three-dimensional adhesive network traps<br />

in the presence of nematodes (Nordbring-Hertz 1977). Apart from a low


194 L.V. Lopez-Llorca et al.<br />

nutrient status, small peptides, e.g. phenylalanyl-valine, can induce trap<br />

formation (Nordbring-Hertz 1973). When traps are formed, and even before<br />

traps are fully developed, nematodes can be captured in the adhesive<br />

covering the traps. The adhesive has been partially characterised and appears<br />

to be a polymer complex containing proteins, neutral sugars and<br />

uronic acids (Tunlid et al. 1991). The adhesive changes properties, from<br />

an amorphous stage to directed fibrils, after contact with the nematode<br />

cuticle (Veenhuis et al. 1985). This is in contrast to the endoparasitic fungus<br />

Drechmeria coniospora, wherethefibrilsappeardirectedwhethernematodes<br />

are present or not (Jansson and Nordbring-Hertz 1988). After<br />

adhesion, the fungus penetrates the nematode cuticle from the trap, probably<br />

using both mechanical and enzymatic means. Since the nematode<br />

cuticle contains mainly proteinaceous material (Bird and Bird 1991), extracellular<br />

proteolytic enzymes involved in cuticle penetration have been<br />

studied. The major protease appears to be subtilisin PII. This serine protease<br />

has been characterised and genomically cloned (Åhman et al. 1996).<br />

After penetration, an infection bulb is formed, from which trophic hyphae<br />

grow out and digest the contents of the nematode. New hyphae and traps<br />

arethenformedoutsidethenematodecorpustostartanewinfection<br />

cycle.<br />

Pochonia spp. adhere to nematode egg shells by means of an appressorium<br />

formed at the hyphal tip (Lopez-Llorca and Claugher 1990, Lopez-<br />

Llorca et al. 2002b). An extracellular material (ECM) probably functions as<br />

adhesive, but possibly also seals the perforation in the egg shell caused by<br />

the penetration hypha beneath the fungal appressorium. This extracellular<br />

material can be labelled with the lectin Concanavalin A, indicating that<br />

the ECM contains mannose/glucose moieties probably on the side chains<br />

of glycoproteins (Lopez-Llorca et al. 2002b). Most ECMs of fungal hyphae<br />

consist of proteins and carbohydrates (Nicholson 1996). The nematode egg<br />

shell consists mainly of proteins and chitin (Bird and Bird 1991) and therefore<br />

proteases and chitinases would be important for fungal penetration of<br />

the egg shell. Serine proteases have been isolated from Pochonia rubescens,<br />

P32 (Lopez-Llorca 1990), P. chlamydosporia, VcP1 (Segers et al. 1994) and<br />

Paecilomyces lilacinus, PL (Bonants et al. 1995). In some cases these have<br />

been immunolocalised in infected hosts, their peptides sequenced, and the<br />

coding genes cloned. Recently, the chitinolytic system of P. rubescens and<br />

P. chlamydosporia has been studied (Tikhonov et al. 2002). For both species,<br />

among other enzymes, a similar major 43 kDa endochitinase (CHI43) was<br />

purified and characterised. A combination of protease P32 and chitinase<br />

CHI43, or the enzymes individually, caused removal of egg shell layers of<br />

the potato cyst nematode Globodera pallida (Tikhonov et al. 2002). Following<br />

penetration of the egg shell, the fungus digests the contents of<br />

the egg, proliferates inside and later grows outside the egg to penetrate


11 Nematophagous Fungi as Root Endophytes 195<br />

neighbouring eggs in nematode cysts or egg masses; alternatively it can<br />

grow saprophytically.<br />

11.2.2<br />

Mycoparasites<br />

Mycoparasitism is a common feature of fungi (Jeffries 1997). The ability<br />

of nematophagous fungi to attack other fungi was first described by Tzean<br />

and Estey (1978). Nematode-trapping fungi such as A. oligospora attack<br />

their host fungi, e.g. Rhizoctonia solani, in a manner similar to that of the<br />

well known mycoparasite Trichoderma spp. (Chet et al. 1981). The mycoparasitic<br />

behaviour of A. oligospora takes place by coiling of the hyphae of<br />

the nematode-trapping fungi around the host hyphae, which, in contrast to<br />

Trichoderma spp., results in disintegration of the host cell cytoplasm without<br />

penetration of the host (Persson et al. 1985). It has been shown using<br />

radioactive phosphorous tracing that nutrient transfer takes place between<br />

the nematode-trapping fungus A. oligospora and its host R. solani (Olsson<br />

and Persson 1994). Although this phenomenon has never been observed in<br />

soil, it may increase the fitness of the nematode-trapping fungi in soil by<br />

reducing competition and providing nutrients. Moreover, it may extend the<br />

biocontrol capability of nematophagous fungi as biocontrol agents to fungal<br />

parasites as well as nematodes. Furthermore, P. chlamydosporia has been<br />

described as being able to infect propagules of important plant pathogens,<br />

such as uredospores of rust fungi (Leinhos and Buchenauer 1992), and<br />

oospores of Phytophthora and other Oomycetes (Sneh et al. 1977).<br />

11.2.3<br />

Root Endophytes<br />

Most work on the root biology of nematophagous fungi has concerned<br />

externalrootcolonisation(ectorhizosphere).Lately,colonisationintheroot<br />

tissues has also been reported (endorhizosphere). Some of these studies<br />

will be discussed in this chapter.<br />

Ectorhizosphere<br />

Since plant-parasitic nematodes generally attack plant roots it has been<br />

an important task to study the rhizosphere biology of nematophagous<br />

fungi – the root zone is an area with an abundant supply of the nematode<br />

prey. Not surprisingly, the nematode-trapping fungi have been found to<br />

be more frequent in the rhizosphere than in the bulk soil (Peterson and<br />

Katznelson 1965; Gaspard and Mankau 1986; Persmark and Jansson 1997).


196 L.V. Lopez-Llorca et al.<br />

Egg-parasitic fungi were also found to be more abundant in the rhizosphere<br />

(Bourne et al. 1996; Kerry 2000).<br />

External root colonisation varies between plant species. For instance,<br />

Persmark and Jansson (1997) studied the presence and frequency of nematode-trapping<br />

fungi in field soils planted with barley, pea or white mustard.<br />

The pea rhizosphere harboured by far the highest frequency of nematodetrapping<br />

fungi: 19 times higher than in the root free soil. The number<br />

of species of nematode-trapping fungi was also higher in the pea rhizosphere,<br />

with A. oligospora as being the most common species (Persmark<br />

and Jansson 1997). In an investigation on chemotropic growth towards<br />

roots of pea, barley and white mustard by seven species of nematophagous<br />

fungi, only isolates of A. oligospora were attracted to the roots of all plants,<br />

but this was confined to the 2 mm closest to the roots (Bordallo et al.<br />

2002). In a pot experiment, the colonisation of tomato roots by several nematophagousfungiwasfollowedfor3months.Monacrosporiumellipsosporum<br />

and Arthrobotrys dactyloides were especially competent in colonising<br />

the roots (Persson and Jansson 1999). Several nematode-trapping fungi are<br />

able to form so-called “conidial traps” in response to roots and root exudates<br />

(Persmark and Nordbring-Hertz 1997). The conidial traps are capture<br />

organs formed directly on the conidia without hyphal growth, and give the<br />

fungi an extra advantage by spreading the fungi and capturing nematodes,<br />

similar to most endoparasitic fungi, which infect nematodes entirely with<br />

adhesive conidia.<br />

External root colonisation by the egg-parasite Pochonia chlamydosporia<br />

also varied with plant species. For instance, kale and cabbage had a density<br />

of the fungus twice as high as that on soya bean and tomato (Bourne<br />

et al. 1996). Furthermore, rhizosphere colonisation by P. chlamydosporia<br />

was increased when plants were infected with the root-knot nematode<br />

Meloidogyne incognita. Thiseffectispossiblyduetoincreasedleakageof<br />

root exudates after damage to the root surface by the nematodes (Bourne<br />

et al. 1996). In none of the investigations mentioned above was the fungal<br />

colonisation of internal root tissues examined.<br />

Endorhizosphere<br />

Using axenic barley and tomato plants inoculated with the nematophagous<br />

fungi P. chlamydosporia or A. oligospora, we found that both fungi have the<br />

capacity to colonise epidermis and root cortex of barley and the epidermis<br />

of tomato (Lopez-Llorca et al. 2002a, Bordallo et al. 2002).<br />

In these experiments roots were sequentially sampled, cryo-sectioned,<br />

and observed under light- or cryo-scanning electron microscopes. Both<br />

fungi grew inter- and intra-cellularly and formed appressoria (Figs. 11.2a,b)<br />

whenpenetratingplantcellwallsofepidermisandcortexcells,butnever


11 Nematophagous Fungi as Root Endophytes 197<br />

Fig.11.2. a Formation of appressoria (arrow)byArthrobotrys oligospora during penetration<br />

of the epidermis of barley roots. (From Bordallo et al. 2002, courtesy of Blackwell Science).<br />

b Colonisation of tomato roots by Pochonia chlamydosporia. Note appressorium and cell<br />

wall protein apposition (arrow). (From Bordallo et al. 2002, courtesy of Blackwell Science).<br />

c Colonisation of barley roots by A. oligospora. Callose deposit in papillae (arrow) stained<br />

with Sirofluor. (From Bordallo et al. 2002, courtesy of Blackwell Science). d Colonisation<br />

of tomato roots by Pochonia chlamydosporia. Note externally produced chlamydospores<br />

(arrow). (From Bordallo et al. 2002, courtesy of Blackwell Science). Bars a, c 20 µm; b,<br />

d 30 µm<br />

entered vascular tissues. In contrast to Pochonia spp., appressoria had<br />

previously never been observed in A. oligospora. Using histochemical<br />

stains, we could show plant defence reactions, e.g. papillae, lignitubers<br />

and other cell wall appositions induced by nematophagous fungi, but these<br />

never prevented root colonisation. Callose deposits in papillae induced by<br />

A. oligospora in barley roots are shown in Fig. 11.2c. Nematophagous fungi<br />

grew extensively, especially in monocotyledonous plants, producing abundant<br />

mycelia, conidia and chlamydospores (P. chlamydosporia) (Fig. 11.2d).<br />

Necrotic areas of the roots were observed at initial stages of colonisation<br />

by A. oligospora, but were never seen at later stages even when the fungus<br />

proliferated in epidermal and cortical cells. Roots colonised by P. chlamydosporia<br />

displayed higher proteolytic activity than non-inoculated control<br />

roots using immunochemical techniques. The significance of this fact for<br />

biological control of root pathogens is under investigation in our laboratory.


198 L.V. Lopez-Llorca et al.<br />

The growth of the two nematophagous fungi in plant roots appears to resemble<br />

that of an endophyte, i.e. the host remains asymptomatic. Whether<br />

this endophytic growth induces systemic resistance to nematodes and/or<br />

plant pathogens in plants is as yet unknown, but worth further investigation.<br />

We have found that P. chlamydosporia could reduce growth of the<br />

plant-pathogenic fungus Gaeumannomyces graminis var. tritici (take-all<br />

fungus, Ggt) in dual culture Petri dish and in growth tube experiments. In<br />

pot experiments, P. chlamydosporia increased plant growth whether Ggt<br />

was present in the roots or not, suggesting a growth promoting effect by<br />

P. chlamydosporia (Monfort et al. 2005), as has also been found in the case<br />

of colonisation by other endophytic fungi (see Chap. 15 by Schulz).<br />

In a recent screening in our laboratory on the capacity of various types<br />

of nematophagous fungi to grow endophytically in roots, we have shown<br />

that fungi other than A. oligospora and P. chlamydosporia had this ability<br />

(Table 11.1). With these fungi, all four ecological groups of nematophagous<br />

fungi are represented. The results regarding root colonisation are shown in<br />

Table 11.2.<br />

Hirsutella rhossiliensis, whichinfectsnematodesbymeansofadhesive<br />

conidia (Jaffee and Zehr 1982), behaves ecologically as an endoparasitic<br />

fungus (“obligate” parasite), although it can grow in the laboratory on<br />

artificial media. The fungus reacts in a density-dependent manner (Jaffee<br />

et al. 1992) and, in spite of its low efficiency as a nematode antagonist, it<br />

suppresses plant-parasitic nematodes in agroecosystems with little human<br />

disturbance, such as old peach orchards in the United States (Stirling 1991).<br />

H. rhossiliensis, unlike A. oligospora and P. chlamydosporia,doesnotseem<br />

to colonise barley roots endophytically. Three weeks after inoculation,<br />

cortex and epidermal cells were free from hyphal colonisation (Table 11.2).<br />

However, the fungus seems to colonise the rhizoplane abundantly, where it<br />

forms viable conidiophores (Fig. 11.3a).<br />

Nematoctonus (teleomorphHohenbuehelia)isapeculiargenusofbasidiomycetous<br />

nematophagous fungi. It comprises about 15 species, some of<br />

Table 11.1. Endophytic root colonisation of barley by the four ecological groups of nematophagous<br />

fungi<br />

Fungus Type Colonisation Reference<br />

Pochonia chlamydosporia Egg-parasite + Lopez-Llorca et al. 2002a<br />

Arthrobotrys oligospora Nematode-trapping + Bordallo et al. 2002<br />

Arthrobotrys dactyloides Nematode-trapping + This chapter<br />

Nematoctonus robustus Nematode-trapping + This chapter<br />

Nematoctonus pachysporus Endoparasite – This chapter<br />

Hirsutella rhossiliensis Endoparasite – This chapter<br />

Pleurotus djamor Toxin producing + This chapter


11 Nematophagous Fungi as Root Endophytes 199<br />

Table 11.2. Semiquantitative endophytic barley root colonisation by nematophagous fungi<br />

(unpublished results). In these experiments barley seeds were sterilised, germinated and<br />

planted together with the appropriate fungus in culture tubes containing sterilised vermiculite<br />

and water according to Bordallo et al. (2002). After 21 days the roots were harvested,<br />

surface sterilised to remove external hyphal growth, cut into 1-cm fragments and plated on<br />

corn meal agar. Half of the roots were not surface sterilised. After approx 7 days the root<br />

fragments were examined and fungal growth was recorded. Data are based on three roots,<br />

with six fragments of each (n = 18)<br />

Fungus Type Root fragments with fungus (%)<br />

Non-surface Surface<br />

sterilised sterilised<br />

Arthrobotrys dactyloides Nematode-trapping 77.8 5.6<br />

Nematoctonus robustus Nematode-trapping 100.0 22.0<br />

Nematoctonus pachysporus Endoparasite 50.0 0<br />

Hirsutella rhossiliensis Endoparasite 100.0 0<br />

Pleurotus djamor Toxin producing 77.8 11.1<br />

which fit the endoparasitic behaviour, infecting nematodes by means of<br />

adhesive conidia, whereas others capture nematodes with adhesive traps,<br />

yet others share both types of behaviour (Thorn and Barron 1986). This<br />

diverse nematophagous behaviour was also reflected in root colonisation<br />

by the two species studied. Whereas the nematode-trapping N. robustus<br />

penetrated and colonised barley roots (Fig. 11.3b) and formed typical<br />

clamp-connections (Fig. 11.3c) as soon as 1 week after inoculation, the<br />

endoparasite N. pachysporus did not enter the roots but colonised the root<br />

surfaceasdidH. rhossiliensis (Table 11.2). It thus appears that, in general,<br />

the endoparasitic fungi may not be endophytic root colonisers, but this may<br />

also be a reflection of their slow growth rate and their mode of infecting<br />

nematodes.<br />

The genus Pleurotus also belongs to the Basidiomycetes.Itformsbasidiocarps<br />

and its standard means of living is saprophytic growth on decaying<br />

wood. The most common species, P. ostreatus, is a commercially cultivated<br />

edible mushroom. The fungus compensates for the lack of nitrogen<br />

in wood, its natural substrate, with its nematophagous habit (Thorn and<br />

Barron 1984). In fact, the nematophagous habit has been described for<br />

several species of Pleurotus (Thorn and Barron 1984). The fungus immobilises<br />

nematodes with a toxin (Kwok et al. 1992) prior to infection and<br />

digestion of its prey (Nordbring-Hertz et al. 1995). The strong relation with<br />

plant tissues, as a natural wood decomposer, was also confirmed in roots,<br />

when we found that, just as N. robustus, P. djamor penetrated early, and<br />

extensively colonised barley roots (Fig. 11.3d). The fungus was in fact very<br />

aggressive since some parts of the root appeared to be decorticated 3 weeks


200 L.V. Lopez-Llorca et al.<br />

Fig.11.3. a Rhizoplane colonisation of barley roots by the endoparasitic fungus Hirsutella<br />

rhossiliensis 3 weeks after inoculation, forming conidiophore with phialide (arrow).<br />

b Colonisation of cortex of barley roots by the nematode-trapping basidiomycete Nematoctonus<br />

robustus 2 weeks after inoculation. c Colonisation of cortex and epidermis of barley<br />

roots by the nematode-trapping basidiomycete Nematoctonus robustus 1 week after inoculation.<br />

Note clamp-connection (arrow). d Colonisation of cortex of barley roots by Pleurotus<br />

djamor 10 days after inoculation. e Colonisation of cortex of barley roots by the nematodetrapping<br />

fungus Arthrobotrys dactyloides 2 weeks after inoculation. f Colonisation of cortex<br />

of barley roots by the nematode-trapping fungus Arthrobotrys dactyloides 2weeksafter<br />

inoculation showing coiling structure. Bars a, b 30 µm; c–f 15 µm<br />

after inoculation, as was A. oligospora in previous experiments (Bordallo<br />

et al. 2002).<br />

Arthrobotrys dactyloides is a nematode-trapping fungus that captures<br />

nematodes by means of constricting rings (Barron 1977). The fungus was<br />

shown to form functional traps in soil (Jansson et al. 2000) and on the surface<br />

of tomato roots infected with root-knot nematodes (Riekert and Tiedt


11 Nematophagous Fungi as Root Endophytes 201<br />

1994), and has also been used in biological control experiments of plant<br />

parasitic nematodes (Stirling and Smith 1998). In our recent experiments,<br />

A. dactyloides was an active, and early, root coloniser. We found evidence<br />

of epidermal cell penetration and colonisation (Fig. 11.3e) 1 week after<br />

inoculation. Like P. chlamydosporia, the fungus formed coiling structures<br />

in barley root cells (Fig. 11.3f) and extensively colonised the roots. Such<br />

structures are also formed by other root endophytes, e.g. Piriformospora<br />

indica (Varma et al. 1999; Chap. 15 by Schulz), and presumably improve<br />

the exchange of metabolites.<br />

These preliminary studies lack the most important component, the nematode.<br />

With trapping and endoparasitic nematophagous fungi, such experiments<br />

are easy to perform axenically, since the nematophagous habit<br />

can easily be triggered by free-living nematodes. Such experiments are<br />

underway in our laboratory to address the question whether the root<br />

endophytic behaviour of nematophagous fungi is functional in their nematophagous<br />

habit, i.e. if the mycelium growing in roots is able to develop<br />

active trapping organs or adhesive spores.<br />

In the case of fungal nematode egg parasites the inclusion of a plant<br />

parasitic nematode is technically more difficult. This is because the natural<br />

targets of nematophagous Pochonia spp. are cyst and root knot nematodes.<br />

These phytopathogenic nematodes have an endoparasitic behaviour<br />

and a life span of at least a month – too long for our axenic system. Although<br />

these nematodes can be multiplied in axenic systems based on<br />

Agrobacterium-transformed roots on a tissue culture medium (Verdejo-<br />

Lucas 1995), these extremely rich conditions are incompatible for studying<br />

root colonisation by nematophagous fungi. An alternative, which we have<br />

explored (unpublished) is the use of migratory endoparasitic nematodes<br />

such as Pratylenchus spp., some species of which have broad host specificity<br />

and can infect cereals such as barley under our conditions. We have<br />

encountered two problems that have prevented further development. One<br />

is that Pratylenchus spp. is not a host of Pochonia spp. (or has not been<br />

describedyet).Sincethenematodelayseggsinroottissue,onecouldexpect<br />

to find egg infection. However, in our experiments we found that the axenic<br />

conditions are too favourable for the nematode, which multiplies much<br />

faster than the fungus. This may be a question of optimising inoculum. We<br />

are trying to adapt our methods to include endoparasitic nematodes such<br />

as Meloidogyne spp.,andthefinalgoalofourstudiesistodevelopbiological<br />

control strategies to such severe plant pathogens. Development of<br />

alternative control methods, e.g. biological control, is especially important<br />

since the phasing out of the most widespread method of plant parasitic<br />

nematode control, fumigation with methyl bromide, is to be enforced.<br />

Endophytic rhizobacteria that reduce plant-parasitic nematodes have<br />

also been described [Hallman et al. 2001; Chaps. 4 (Berg and Hallmann)


202 L.V. Lopez-Llorca et al.<br />

and 3 (Kloepper and Ryu)], as well as arbuscular mycorrhizal fungi that<br />

reduce root knot nematodes (Waecke et al. 2001). If this is also the case for<br />

nematophagous fungi this will open up a new area of biocontrol using these<br />

fungi. The internal root colonisation by egg-parasitic fungi, e.g. Pochonia<br />

spp.,maygivethefungianopportunitytoinfectnematodeeggsineggsacks<br />

of root-knot nematodes inside the roots and reduce subsequent spread and<br />

infection of roots by the second stage juveniles. Structures resembling trapping<br />

organs were observed in epidermal cells colonized by A. oligospora<br />

(Bordallo et al. 2002), and these may serve the purpose of trapping newly<br />

hatched juveniles escaping the roots. The ability to colonise plant roots<br />

may also be a survival strategy of these fungi and could explain soil suppressiveness<br />

to plant-parasitic nematodes in nature. Root inoculation of<br />

nematophagous fungi may help us to circumvent the lack of receptivity<br />

of soil (even sand) to inoculum of nematophagous fungi (Monfort et al.<br />

2006), which has hampered the capability of fungi such as P. chlamydosporia<br />

to control agronomically important nematodes such as Meloidogyne<br />

spp. (Verdejo-Lucas et al. 2003). This, despite the fact that the nematode is<br />

naturally found to be infected by P. chlamydosporia and other egg-parasitic<br />

fungi in similar Mediterranean agroecosystems (Verdejo-Lucas et al. 2002;<br />

Olivares-Bernabeu and Lopez-Llorca 2002). The root colonisation of plant<br />

roots is a new area of research that deserves in-depth investigation, particularly<br />

for biocontrol purposes.<br />

11.3<br />

Concluding Remarks<br />

The role of the host plant in the tritrophic relationship between nematophagous<br />

fungi, plants and phytopathogenic nematodes has largely been<br />

neglected. In this chapter we have collected and drawn conclusions from<br />

the data accumulated in the literature. We have also presented our own data,<br />

which indicate that the outcome of the interaction between nematophagous<br />

fungi and plants depends both on the host plant (mono- vs di-cotyledon)<br />

and the fungal species, but also on the ecological groups of nematophagous<br />

fungi (trapping, endoparasitic, egg-parasite, toxin producer).<br />

The plant host is, after all, the most important living entity in an agroecosystem<br />

and is, of course, the target of any approach to disease control.<br />

We would like to stress the biological – theoretical – importance of the<br />

endophytic behaviour of nematophagous fungi, for instance, as a likely<br />

means to explain soil suppressiveness to plant parasitic nematodes. Another<br />

theoretical component is the possible discovery of a new mechanism<br />

of the mode of action of nematophagous fungi, that of interaction with<br />

the plant host. This may function in two ways. We have initial evidence of


11 Nematophagous Fungi as Root Endophytes 203<br />

plant growth promotion. The other function could be modulation of plant<br />

defences. Similar activities have been found for antagonists (bacterial and<br />

fungal) of other plant pathogens, but also for other endophytic fungi that<br />

are not necessarily antagonists [see Chaps. 3 (Kloepper and Ryu), 4 (Berg<br />

and Hallmann) and 14 (Schulz)].<br />

Basic discoveries in the field of pathogen–plant host interaction have<br />

led to new developments and approaches to plant disease control. For<br />

instance, the recent proliferation of compounds modulating plant defence<br />

responses as “magic bullets” to control a wide array of biotic and abiotic<br />

stresses in plants had its origin in the study of the molecular basis of<br />

virulence/avirulence in the plant pathogenic bacteria–plant interaction.<br />

There are similar examples, such as the effect of oligosaccharides or other<br />

molecules on plant defence systems. A future possibility for enhancing<br />

control of plant pathogens may be to devise a way to obtain synergies with<br />

non-chemical means of control. Perhaps we are in the dawn of a new era of<br />

biological control of nematodes that may circumvent the difficulties found<br />

in the mere application of nematophagous fungi to soil.<br />

<strong>References</strong><br />

Åhman J, Ek B, Rask L, Tunlid A (1996) Sequence analysis and regulation of a gene encoding<br />

a cuticle-degrading serine protease from the nematophagous fungus Arthrobotrys<br />

oligospora. Microbiology 142:1605–1616<br />

Ahrén D, Ursing BM, Tunlid A (1998) Phylogeny of nematode trapping fungi based on 18S<br />

rDNA sequences. FEMS Microbiol Lett 158:179–184<br />

Barron GL (1977) The nematode-destroying fungi. Topics in mycobiology No. 1 Canadian<br />

Biological Publications, Guelph<br />

Bird AF, Bird J (1991) The structure of nematodes. Academic, San Diego<br />

Bonants PJM, Fitters PFL, Thijs H, Den Belder E, Waalwijk C, Willem J, Henfling DM (1995)<br />

A basic serine protease from Paecilomyces lilacinus with biological activity against<br />

Meloidogyne hapla eggs. Microbiology 141:775–784<br />

Bordallo JJ, Lopez-Llorca LV, Jansson H-B, Salinas J, Persmark L, Asensio L (2002) Effects<br />

of egg-parasitic and nematode-trapping fungi on plant roots. New Phytol 154:491–499<br />

Bourne JM, Kerry BR, De Leij FAAM (1996) The importance of the host plant<br />

on the interaction between root-knot nematodes (Meloidogyne spp.)andthenematophagous<br />

fungus, Verticillium chlamydosporium Goddard. Biocontrol Sci Technol<br />

6:539–548<br />

Chet I, Harman GER, Baker R (1981) Trichoderma hamatum: its hyphal interaction with<br />

Rhizoctonia solani and Pythium spp. Microb Ecol 7:29–38<br />

Dackman C, Jansson H-B, Nordbring-Hertz B (1992) Nematophagous fungi and their activities<br />

in soil. In: Bollag J-M, Stotzky G (eds) Soil biochemistry, vol 7 Decker, New York,<br />

pp 95–130<br />

Gams W, Zare R (2003) A taxonomic review of the clavicipitaceous anamorphs parasitizing<br />

nematodes and other microinvertebrates. In: White JF, Bacon CW, Hywel-Jones NL,<br />

Spatafora JW (eds). Clavicipitalean fungi: evolutionary biology, chemistry, biocontrol,<br />

and cultural impacts Dekker, New York, pp 17–73


204 L.V. Lopez-Llorca et al.<br />

Gaspard JT, Mankau R (1986) Nematophagous fungi associated with Tylenchulus semipenetrans<br />

and the citrus rhizosphere. Nematologica 32:359–363<br />

Hagedorn G, Scholler M (1999) A reevaluation of predatory orbiliaceous fungi. I. Phylogenetic<br />

analysis using rDNA sequence data. Sydowia 51:27–48<br />

Hallmann J, Quadt-Hallmann A, Miller WG, Sikora RA, Lindow SE (2001) Endophytic colonization<br />

of plants by the biocontrol agent Rhizobium etli G12 in relation to Meloidogyne<br />

incognita infection. Phytopathology 91:415–422<br />

Jaffee BA, Zehr EI (1982) Parasitism of the nematode Criconemella xenoplax by the fungus<br />

Hirsutella rhossiliensis. Phytopathology 72:1378–1381<br />

Jaffee B, Phillips R, Muldoon A, Mangel M (1992) Density-dependent host pathogen dynamics<br />

in soil microcosms. Ecology 73:495–506<br />

Jansson H-B, Lopez-Llorca LV (2001) Biology of nematophagous fungi. In: Misra JK,<br />

Horn BW (eds). Trichomycetes and other fungal groups. Science Publishers, Enfield,<br />

pp 145–173<br />

Jansson H-B, Lopez-Llorca LV (2004) Control of nematodes by fungi. In: Arora DK (ed)<br />

Fungal biotechnology in agriculture, food, and environmental applications Dekker,<br />

New York, pp 205–215<br />

Jansson H-B, Nordbring-Hertz B (1983) The endoparasitic nematophagous fungi Meria<br />

coniospora infects nematodes specifically at the chemosensory organs. J Gen Microbiol<br />

129:1121–1126<br />

Jansson H-B, Nordbring-Hertz B (1988) Infection events in the fungus-nematode system.<br />

In: Poinar GO, Jansson H-B (eds) Diseases of nematodes, vol 2. CRC, Boca Raton,<br />

pp 59–72<br />

Jansson H-B, Persson C, Odselius R (2000) Growth and capture activities of nematophagous<br />

fungi in soil visualized by low temperature scanning electron microscopy. Mycologia<br />

92:10–15<br />

Jeffries P (1997) Mycoparasitism. In: Dicklow TD, Söderström B (eds). The mycota IV. Environmental<br />

and microbiological relationships, Springer, Berlin Heidelberg New York,<br />

pp 149–164<br />

Kerry BR (2000) Rhizosphere interactions and the exploitation of microbial agents for the<br />

biological control of plant-parasitic nematodes. Annu Rev Phytopathol 38:423–441<br />

Kwok OCH, Plattner R, Weisleder D, Wicklow DT (1992) A nematicidal toxin from Pleurotus<br />

ostreatus NRRL 3526. J Chem Ecol 18:127–136<br />

Leinhos GME, Buchenauer H (1992) Hyperparasitism of selected fungi on rust fungi on<br />

cereals. Z Pflanzenkr Pflanzenschutz 99:482–498<br />

Lopez-Llorca LV (1990) Purification and properties of extracellular proteases produced by<br />

the nematophagous fungus Verticillium suchlasporium. Can J Microbiol 36:530–537<br />

Lopez-Llorca LV, Claugher D (1990) Appressoria of the nematophagous fungus Verticillium<br />

suchlasporium. Micron Microsc Acta 21:125–130<br />

Lopez-Llorca LV, Bordallo JJ, Salinas J, Monfort E, Lopez-Serna ML (2002a) Use of light<br />

and scanning electron microscopy to examine colonisation of barley rhizosphere by the<br />

nematophagous fungus Verticillium chlamydosporium. Micron 33:61–67<br />

Lopez-Llorca LV, Olivares-Bernabeu C, Salinas J, Jansson H-B (2002b) Pre-penetration events<br />

in fungal parasites of nematode eggs. Mycol Res 106:499–506<br />

Meyer SLF (1998) Evaluation of Verticillium lecanii strains applied in root drenches for<br />

suppression of Meloidogyne incognita on tomato. J Helminthol 65:82–86<br />

Monfort E, Lopez-Llorca LV, Jansson H-B, Salinas J, Park J-O, Sivasithamparam K (2005)<br />

Colonisation of seminal roots of wheat and barley by egg-parasitic nematophagous<br />

fungi and their effects on Gaeumannomyces graminis var. tritici and development of<br />

root-rot. Soil Biol Biochem 37:1229–1235


11 Nematophagous Fungi as Root Endophytes 205<br />

Monfort E, Lopez-Llorca LV, Jansson H-B, Salinas J (2006) In vitro soil receptivity assays to<br />

egg-parasitic nematophagous fungi. Mycol Prog 5:18–23<br />

Nicholson RL (1996) Adhesion of fungal propagules. In: Nicole M, Gianinazzi-Pearson V<br />

(eds) Histology, ultrastructure and molecular cytology of plant-microorganism interactions.<br />

Kluwer, Amsterdam, pp 117–134<br />

Nordbring-Hertz B (1973) Peptide-induced morphogenesis in the predacious fungus<br />

Arthrobotrys oligospora. Physiol Plant 29:223–233<br />

Nordbring-Hertz B (1977) Nematode-induced morphogenesis in the predacious fungus<br />

Arthrobotrys oligospora. Nematologica 23:443–451<br />

Nordbring-Hertz B, Jansson H-B, Friman E, Persson Y, Dackman C, Hard T, Poloczek E,<br />

Feldman R (1995) Nematophagous Fungi Institut für den Wissenschaftlichen Film,<br />

Göttingen. Film No C 1851<br />

Olivares-Bernabeu C, Lopez-Llorca LV (2002) Fungal egg-parasites of plant-parasitic nematodes<br />

from Spanish soils. Rev Iberoam Micol 19:104–110<br />

Olsson S, Persson Y (1994) Transfer of phosphorus from Rhizoctonia solani to the mycoparasite<br />

Arthrobotrys oligospora. Mycol Res 98:1065–1068<br />

Persmark L, Jansson H-B (1997) Nematophagous fungi in the rhizosphere of agricultural<br />

crops. FEMS Microbiol Ecol 22:303–312<br />

Persmark L, Nordbring-Hertz B (1997) Conidial trap formation of nematode-trapping fungi<br />

in soil and soil extracts. FEMS Microbiol Ecol 22:313–323<br />

Persson C, Jansson H-B (1999) Rhizosphere colonization and control of Meloidogyne spp.<br />

by nematode-trapping fungi. J Nematol 31:164–171<br />

Persson Y, Veenhuis M, Nordbring-Hertz B (1985) Morphogenesis and significance of hyphal<br />

coiling by nematode-trapping fungi in mycoparasitic relationships. FEMS Microbiol<br />

Ecol 31:283–291<br />

Peterson EA, Katznelson H (1965) Studies on the relationships between nematodes and<br />

other soil microorganisms. IV. Incidence of nematode-trapping fungi in the vicinity of<br />

plant roots. Can J Microbiol 11:491–495<br />

Pfister DH (1997) Castor, Pollux and life histories of fungi. Mycologia 89:1–23<br />

Poinar GO (1983) The natural history of nematodes. Prentice-Hall, Englewood Cliffs<br />

Riekert HF, Tiedt LR (1994) Scanning electron microscopy of Meloidogyne incognita juveniles<br />

entrapped in maize roots by a nematode-trapping fungus Arthrobotrys dactyloides.<br />

S Afr J Zool 29:189–191<br />

Rubner A (1996) Revision of predacious hyphomycetes in the Dactylella-Monacrosporium<br />

complex. Stud Mycol 39:1–134<br />

Segers R, Butt TM, Kerry BR, Peberdy F (1994) The nematophagous fungus Verticillium<br />

chlamydosporium Goddard produces a chymoelastase-like protease which hydrolyses<br />

host nematode proteins in situ. Microbiology 140:2715–2723<br />

Sneh B, Humble SJ, Lockwood JL (1977) Parasitism of oospores of Phytophthora megasperma<br />

var. sojae, P. cactorum, Pythium and Aphanomyces euteiches by oomycetes,<br />

chytridomycetes, hyphomycetes, actinomycetes and bacteria. Phytopathology<br />

67:622–628<br />

Stirling GR (1991) Biological control of plant parasitic nematodes. Progress, problems and<br />

prospects. CAB International, Wallingford<br />

Stirling GR, Smith LJ (1998) Field tests of formulated products containing either Verticillium<br />

chlamydosporium or Arthrobotrys dactyloides for biological control of root-knot<br />

nematodes. Biol Contr 11:231–239<br />

Thorn RG, Barron GL (1984) Carnivorous mushrooms. Science 224:76–78<br />

Thorn RG, Barron GL (1986) Nematoctonus and the Tribe resupinatae in Ontario, Canada.<br />

Mycotaxon 25:321–453


206 L.V. Lopez-Llorca et al.<br />

Tikhonov VE, Lopez-Llorca LV, Salinas J, Jansson H-B (2002) Purification and characterization<br />

of chitinases from the nematophagous fungi Verticillium chlamydosporium and<br />

V. suchlasporium. Fungal Gen Biol 35:67–78<br />

Tunlid A, Johansson B, Nordbring-Hertz B (1991) Surface polymers of the nematodetrapping<br />

fungus Arthrobotrys oligospora. J Gen Microbiol 137:1231–1240<br />

Tzean SS, Estey RH (1978) Nematode-trapping fungi as mycopathogens. Phytopathology<br />

68:1266–1270<br />

Varma A, Verma S, Sudha, Sahay N, Bütehorn B, Franken P (1999) Piriformospora indica,<br />

a cultivable plant-growth-promoting root endophyte. Appl Environ Microbiol 65:2741–<br />

2744<br />

Veenhuis M, Nordbring-Hertz B, Harder W (1985) An electron-microscopical analysis of<br />

capture and initial stages of penetration of nematodes by Arthrobotrys oligospora.Antonie<br />

van Leeuwenhoek 51:385–398<br />

Verdejo-Lucas S (1995) Dual cultures: nematodes. In: Singh RP, Singh US (eds) Molecular<br />

methods in plant pathology. CRC, Boca Raton, pp 301–312<br />

Verdejo-Lucas S, Ornat C, Sorribas FJ, Stchiegel A (2002) Species of root-knot nematodes<br />

and fungal egg parasites recovered from vegetables in Almeria and Barcelona, Spain.<br />

J Nematol 34:405–408<br />

Verdejo-Lucas S, Sorribas FJ, Ornat C, Galeano M (2003) Evaluating Pochonia chlamydosporia<br />

in a double-cropping system of lettuce and tomato in plastic houses infested with<br />

Meloidogyne javanica. Plant Pathol 52:521–528<br />

Waecke JW, Waudo SW, Sikora R (2001) Suppression of Meloidogyne hapla by arbuscular<br />

mycorrhiza fungi (AMF) on pyrethrum in Kenya. Int J Pest Manage 47:135–140


12<br />

12.1<br />

Introduction<br />

Molecular Diversity and Ecological<br />

Roles of Mycorrhiza-Associated Sterile<br />

Fungal Endophytes in Mediterranean<br />

Ecosystems<br />

Mariangela Girlanda, Silvia Perotto, Anna Maria Luppi<br />

Under natural conditions, plant roots sustain considerable fungal diversity<br />

(Vandenkoornhuyse et al. 2002). In healthy plants, colonisation of root<br />

tissuesisnotrestrictedtomycorrhizalsymbionts;otherfungicanalsogrow<br />

asymptomatically in the roots, occurring either in the presence or absence<br />

of ecto- or endo-mycorrhizal mycobionts. Isolation from surface-sterilised<br />

roots usually yields a great proportion of fungi with dematiaceous hyphae<br />

and ascomycetous septa, which are sterile in culture, and are referred to as<br />

“dark septate endophytes” (DSE) or “dark sterile mycelia” (DSM) (Schild<br />

et al. 1988; Summerbell 1989; Read 1991; Holdenrieder and Sieber 1992;<br />

Girlanda and Luppi-Mosca 1995; Ahlich and Sieber 1996; Ahlich et al. 1998;<br />

Jumpponen and Trappe 1998a; Schadt et al. 2001; see Chap. 7 by Sieber and<br />

Grünig). These mycelia can also be directly observed to grow inter- and<br />

intra-cellularly in the root cortex, producing peculiar structures, such as<br />

microsclerotia that fill cortical cells (Jumpponen and Trappe 1998a; Barrow<br />

and Aaltonen 2001; Yu et al. 2001; Kovacs and Szigetvari 2002; Ruotsalainen<br />

et al. 2002; Barrow 2003). By virtue of their constant association with roots,<br />

they can be qualified as true root symbionts (see Chap. 1 by Schulz and<br />

Boyle).<br />

Although the presence of DSE in roots was noted early in the last century<br />

(Melin 1922, 1923; Peyronel 1924), their abundant, regular, and ubiquitous<br />

occurrence was given prominence only recently, and is suggestive of<br />

a significant role in natural ecosystems. As a group, these fungi colonise<br />

a broad range of hosts, being reported from nearly 600 plant species, representing<br />

about 320 genera and 114 families (Jumpponen and Trappe 1998a).<br />

However, DSE represent a heterogeneous assemblage of ascomycetous taxa.<br />

Mariangela Girlanda: Dipartimento di Biologia Vegetale and IPP - Torino, Viale PA Mattioli<br />

25, 10125 Torino, Italy, E-mail: mariangela.girlanda@unito.it<br />

Silvia Perotto: Dipartimento di Biologia Vegetale and IPP - Torino, Viale PA Mattioli 25,<br />

10125 Torino, Italy<br />

Anna Maria Luppi: Dipartimento di Biologia Vegetale, Viale PA Mattioli 25, 10125 Torino,<br />

Italy<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


208 M. Girlanda et al.<br />

Although these fungi are mostly sterile when brought into culture, sporulation<br />

has been occasionally induced under particular incubation conditions<br />

(Richard and Fortin 1973; Wang and Wilcox 1985; Fernando and Currah<br />

1995; Ahlich and Sieber 1996), leading to recognition of distinct conidial<br />

forms (Gams 1963; Deacon 1973; Richard and Fortin 1973; Wang and<br />

Wilcox 1985; Currah et al. 1987). In persistently sterile isolates, which are<br />

unidentifiable with conventional criteria, systematic heterogeneity is indicated<br />

by different cultural macro- and micro-scopic morphologies and by<br />

polymorphisms revealed by molecular markers (Stoyke et al. 1992; Harney<br />

et al. 1997; Jumpponen and Trappe 1998a; Schadt et al. 2001; Addy et al.<br />

2001; Grünig et al. 2001, 2002b; see Chap. 7 by Sieber and Grünig). For<br />

instance, sequencing of the 18S nuclear ribosomal DNA (rDNA) region has<br />

shown polyphyletic placement of DSE fungi within Ascomycetes, with representatives<br />

of distinct orders such as Pleosporales, Leotiales, and Pezizales<br />

(Lobuglio et al. 1996; Jumpponen and Trappe 1998a; Schadt et al. 2001;<br />

Girlanda et al. 2002). As a consequence of such heterogeneity, actual host<br />

specificity might vary within the DSE group. Specificity is an important attribute<br />

of fungus-plant associations, and understanding this aspect of the<br />

association is crucial to clarifying the functional nature of the interactions<br />

with the host plants, and hence the ecological role of DSE associates. Indeed,<br />

inoculation experiments of different plants with different DSE strains<br />

have indicated that plant growth response may depend on the particular<br />

combination of the fungus and plant being tested (Jumpponen and Trappe<br />

1998a; Jumpponen 2001).<br />

To date, most of the knowledge available on the specific identity of DSE<br />

fungi, as assessed by molecular analyses of internal transcribed spacer<br />

(ITS) regions of nuclear rDNA, derives from isolates obtained from subantarctic<br />

and northern temperate forests or from Arctic areas in Europe<br />

and Canada (Stoyke et al. 1992; Harney et al. 1997; Jumpponen 1999; Addy<br />

et al. 2001; Grünig et al. 2001, 2002a, 2002b; Schadt et al. 2001; see Chap. 7<br />

by Sieber and Grünig); other biomes remain largely unexplored in DSE<br />

research. Extending investigations to different environments offers the opportunity<br />

of assessing both host and habitat specificity patterns for these<br />

fungi. Mediterranean ecosystems are especially interesting in this respect.<br />

They develop in temperate-warm climates (mean annual temperature generally<br />

ranging from 14 ◦ to 20 ◦ C), with precipitation of 350–1,000 mm/year,<br />

characterised by hot, dry summers and mild, wet winters (“winter-rain<br />

and summer dry” climates). Such climatic conditions occur in five distinct<br />

regions of the world, i.e. the Mediterranean basin, California, Central Chile,<br />

the Cape Province in South Africa, and Southern Australia (Di Castri and<br />

Mooney 1973). Although other regions may exhibit similar mean annual<br />

temperatures and rainfall, they differ in rain distribution over the year<br />

(Japan, for instance, lacks summer drought). Vegetation in Mediterranean


12 DSE of Mediterranean mycorrhizal plants 209<br />

climate areas has different regional expressions (such as “maquis” and “garrigue”<br />

in the Mediterranean basin, “chaparral” in California, “matorral”<br />

in Chile, “fynbos” and “veld” in capensic flora, “kwongan” and “mallee”<br />

in Australia), which, however, display striking convergence in life forms<br />

as an adaptation to the distinctive climatic regime (Pignatti et al. 2002).<br />

A characteristic, shared feature of Mediterranean vegetation is a high diversity<br />

of plants and their associated mycorrhizal types: sclerophyllous<br />

and evergreen shrubs, and small trees bearing arbuscular, ericoid, arbutoid<br />

mycorrhiza and ectomycorrhiza, which coexist and may take up equal<br />

size and dominance (Allen 1991). These environments offer therefore an<br />

interesting scenario for comparative studies of root-fungus associations in<br />

plants with different mycorrhizal status.<br />

We have been investigating molecular diversity and the possible ecological<br />

roles of DSE associates of host pairs in Mediterranean ecosystems in<br />

Northern Italy (Liguria). Neighbouring, healthy-looking individuals of the<br />

ectomycorrhizal Pinus halepensis Mill. (Halep pine) and Quercus ilex L.<br />

(holm oak) and the endomycorrhizal Rosmarinus officinalis L. (arbuscular<br />

mycorrhiza; rosemary) and Erica arborea L. (ericoid mycorrhiza; Mediterranean<br />

heather) were selected for isolation of DSE fungi. Using both molecular<br />

approaches and synthesis experiments the fungi were characterised<br />

for their range of diversity and for possible ecophysiological traits involved<br />

in interactions with different plant hosts.<br />

12.2<br />

Diversity of DSE Associates of Ecto- and Endo-Mycorrhizal<br />

Plants in Mediterranean Ecosystems in Northern Italy<br />

To assess the diversity of DSE isolates obtained from surface-sterilised mycorrhizal<br />

roots of the two different host pairs (Pinus halepensis / Rosmarinus<br />

officinalis, Quercus ilex / Erica arborea), morphotypes were recognised<br />

based on macro- and micro-scopic somatic features and assessed for consistency<br />

through internal transcribed spacer-restriction fragment length<br />

polymorphism (ITS-RFLP); further molecular characterisation was carried<br />

out on morphotypes repeatedly obtained from both hosts (Girlanda et al.<br />

2002; Bergero et al. 2000, 2003). Such shared DSE groups occurred regularly<br />

and at high frequency. In the P. halepensis / R. officinalis study, for instance,<br />

such groups were isolated in all the collection periods, spanning the course<br />

of 11 years. Ribosomal DNA sequences obtained for representative isolates<br />

from each morphotype were used as queries for BLAST searches in publicDNAdatabasesaswellasforphylogeneticreconstructionscarriedout<br />

on datasets comprising both named alignable sequences retrieved from<br />

BLAST searches and sequences of closely related taxa.


210 M. Girlanda et al.<br />

In the investigation on the P. halepensis / R. officinalis host pair, carried<br />

out at Varigotti (Girlanda et al. 2002), taxonomic affiliation through<br />

sequence analysis of rDNA regions has revealed a peculiar spectrum of<br />

taxa, distinct from those recognised to date in other hosts. In some cases,<br />

ITS (ITS1–5.8S–ITS2) sequence matches with sequences in GenBank were<br />

poor (


12 DSE of Mediterranean mycorrhizal plants 211<br />

(Shoemaker et al. 1990), causing a root and crown rot disease of lucerne<br />

in Australia, characterised by symptoms that are somewhat reminiscent of<br />

those induced by R. vagum on cucurbit hosts (Alcorn and Irwin 1987).<br />

None of the DSE morphotypes investigated, therefore, is identifiable with<br />

any of the taxa known to date as DSE sporulating forms (such as Phialocephala<br />

spp., Phialophora spp., Cadophora finlandia (Wang & Wilcox) Harrington<br />

& McNew, Chloridium paucisporum C.J.K. Wang & H.E. Wilcox,<br />

Leptodontidium orchidicola Sigler & Currah; Gams 1963; Deacon 1973;<br />

Richard and Fortin 1973; Wang and Wilcox 1985; Currah et al. 1987; Harrington<br />

and McNew 2003). Among these, Phialocephala fortinii C.J.K. Wang<br />

& Wilcox appears as the major component of DSE assemblages in northern<br />

alpine and subalpine forest ecosystems, occurring with no apparent host<br />

specificity in North America, Europe, and Japan (Wang and Wilcox 1985;<br />

Currah et al. 1987; Stoyke and Currah 1991, 1993; Stoyke et al. 1992; Currah<br />

and Tsuneda 1993; O’Dell et al. 1993; Ahlich and Sieber 1996; Dahlberg<br />

et al. 1997; Hambleton and Currah 1997; Harney et al. 1997; Addy et al.<br />

2001; Grünig et al. 2001, 2002a, 2002b; see Chap. 7 by Sieber and Grünig).<br />

Although the occurrence of these taxa in the Mediterranean ecosystem we<br />

have studied cannot be ruled out entirely, they certainly do not appear<br />

among the dominant DSE in such an environment, where P. halepensis and<br />

R. officinalis represent the most abundant plant species.<br />

In the Q. ilex (holm oak) /E.arborea(Mediterranean heather) studies<br />

(Bergero et al. 2000, 2003), investigations were carried out at different stages<br />

in the evolution of the plant community at a site near Borgio Verezzi. If<br />

left undisturbed, Mediterranean vegetation in Northern Italy develops into<br />

pure Q. ilex woodland, characterised by an extremely reduced understorey<br />

vegetation, due to the disappearance of several plant species of earlier<br />

stages of succession. Later stages are dominated by ectomycorrhiza, in<br />

contrast to maquis and garrigue vegetation, where ectomycorrhizal plants<br />

often co-dominate with endomycorrhizal hosts, such as E. arborea, atypical<br />

ericoid Mediterranean shrub. In the climax woodland, disturbances<br />

such as human activities and fire, which play a key role in shaping Mediterranean<br />

vegetation, open the way for recolonisation by pioneer plant species,<br />

thus initiating secondary successions. Investigations were carried out both<br />

within a pure woodland where Q. ilex establishment had caused the disappearance<br />

of most pioneer species, including E. arborea, and in post-cutting<br />

clearings where the latter species had become re-established.<br />

Two DSE morphotypes (Sd2 and Sd9) were isolated from both holm oak<br />

and heather roots and characterised in detail. When tested in dual axenic<br />

culture trials on E. arborea, isolates assigned to these morphotypes were<br />

found to be able to form typical intracellular hyphal coils characteristic<br />

of ericoid mycorrhizal infection (see below). Morphotype Sd2 was also<br />

isolated from soil of the thickest part of the mature holm oak woodland


212 M. Girlanda et al.<br />

(under adult trees), using E. arborea as a bait plant. No significant ITS<br />

sequence identity was found in GenBank for this morphotype. However,<br />

based on more conserved sequences such as the 5 ′ end of 28S rDNA, it<br />

was found to cluster with ericoid fungi from Gaultheria shallon in a clade<br />

comprising Capronia species (Chaetothyriomycetidae; Allen et al. 2003).<br />

In contrast, morphotype Sd9 displayed significant matches over the ITS region<br />

with nameless fungi of Helotialean affinities represented in GenBank.<br />

This morphotype apparently belongs to an undefined complex within the<br />

Helotiales, comprising DSE, ericoid endomycorrhizal and ectomycorrhizal<br />

fungi from ericoid (Gaultheria shallon), epacrid (Astroloma pinifolium,<br />

Woollsia pungens), tree and sedge hosts (Tsuga sp., Pinus sylvestris, Betula<br />

pubescens, Kobresia myosuroides) from Australia, Norway, Canada and Colorado<br />

(Monreal et al. 1999; McLean et al. 1999; Vrålstad et al. 2002a; Bergero<br />

et al. 2003; and unpublished sequences). In particular, Sd9 shared high sequence<br />

identity of approx. 99% over ca. 440 bp with an ectomycorrhizal<br />

fungus from the sedge Kobresia myosuroides from Colorado (Fig. 12.1).<br />

Genetic relatedness between sterile mycorrhizal isolates obtained from<br />

ericoid and epacrid hosts in the northern and southern hemispheres, respectively,<br />

has been shown by several authors, in accordance with the<br />

current view of these plants as members of the single family Ericaceae<br />

(Crayn and Quinn 2000), and suggests the monophyletic origin of their<br />

endomycorrhizal associations (see Chap. 14 by Cairney). Such mycorrhizal<br />

fungi of Ericaceae from both hemispheres exhibit phylogenetic affinities to<br />

the Helotiales. Their precise placement within this order, however, remains<br />

unclear: nameless mycorrhizal fungi are either part of a Hymenoscyphus<br />

ericae-relatedgrouporofseparate,unspecifiedgroupsoftenreceivingpoor<br />

bootstrap support (McLean et al. 1999; Monreal et al. 1999; Chambers et<br />

al. 2000; Sharples et al. 2000; Cairney and Ashford 2002; see Chap. 14 by<br />

Cairney). Genetic relatedness has also been demonstrated between Hymenoscyphus<br />

ericae and the fungal symbiont forming Piceirhiza bicolorata<br />

ectomycorrhizae, belonging to a monophyletic aggregate possibly comparable<br />

with a generic unit (Vrålstad et al. 2000). It is thus accepted that<br />

◮ Fig.12.1. Neighbour-joining tree for nuclear rDNA ITS (ITS1–5.8S–ITS2) sequences of<br />

DSE morphotypes Sd1, Sd9, Sm1, Sm2, Sm3, Sm5 and Sm8 and other alignable sequences<br />

from BLAST searches. Morphotypes were isolated from Erica arborea roots (Sd1, Sd9 and<br />

Sm1), Quercus ilex roots (Sd1 and Sd9), pure woodland soil where Q. ilex establishment had<br />

caused the disappearance of E. arborea (Sd1, Sm1, Sm2, Sm5 and Sm8), soil of post-cutting<br />

clearings where the latter species had re-established (Sd1, Sd9, Sm1, Sm2 and Sm3), and<br />

were found to be able to form typical mycorrhizal hyphal coils when tested in dual axenic<br />

culture trials on E. arborea (Bergero et al. 2000, 2003). The Kimura-2-parameter model was<br />

used for pairwise distance measurement. Bootstrap values above 50% are indicated (1,000<br />

replicates). The bar indicates 0.1 bp changes. The tree was rooted automatically


12 DSE of Mediterranean mycorrhizal plants 213


214 M. Girlanda et al.<br />

Helotiales encompass fungi with different strategies of association with<br />

roots, including ecto-, ectoendo-, ericoid endo-mycorrhizal and DSE fungi<br />

(LoBuglio et al. 1996; Monreal et al. 1999; Sharples et al. 2000; Vrålstad et<br />

al. 2000, 2002a). However, subgrouping within the order, and hence the<br />

possible occurrence of strict ericoid-, ecto-, and DSE lineages (Read 2000),<br />

is still unclear, since the boundaries between intra- and inter-specific ITS<br />

variation are presently uncertain for these fungi. More detailed molecular<br />

comparisons, using sequence data from further loci, as well as other genetic<br />

markers are needed before relationships within this group can be resolved<br />

with further detail (Cairney and Ashford 2002). In the case of morphotype<br />

Sd9, however, identical random amplified polymorphic DNA (RAPD) profiles<br />

were obtained for isolates from both E. arborea and Q. ilex, indicating<br />

that not only the same taxon, but also the same genotype of that taxon<br />

is shared between an ericoid and an ectomycorrhizal host (Bergero et al.<br />

2000).<br />

Database searches using ITS sequences from Mediterranean DSEs have<br />

allowed comparisons over geographically distinct areas and even biomes.<br />

Some taxa of Helotialean affinities appear to recur across unrelated biomes,<br />

suggesting that such fungi find a suitable niche within root tissues, independently<br />

of the precise host and environment. The apparent absence<br />

in the Mediterranean habitats investigated of Phialocephala fortinii, the<br />

dominating DSE inhabitant of roots in northern temperate forests, and,<br />

by contrast, the presence of Rhizopycnis vagum (see above) suggests the<br />

existence of some environmental selection on distribution of these fungi.<br />

Within a given biome, while ITS data leave the question of actual host specificity<br />

unresolved, there is evidence from RAPD data on Mediterranean DSE<br />

that some DSE genotypes may actually associate with both ecto- and endomycorrhizal<br />

plants. Such a capacity is a prerequisite for these fungi to play<br />

a role in interactions between different plant hosts.<br />

12.3<br />

Ecological Relationships with Conventional Mycorrhizal<br />

and Pathogenic Symbionts<br />

The functional significance of DSE associations is variously described in<br />

the literature. A range of enzymatic activities has been reported for DSE,<br />

conferring potential ability to utilise some of the major organic detrital<br />

nutrient pools (Bååth and Söderström 1980; Haselwandter 1983; Currah<br />

and Tsuneda 1993; Fernando and Currah 1995; Caldwell et al. 2000). By<br />

using axenically grown Earboreaas a bait plant, we have shown that soil<br />

from mature Qilexforest from which the ericoid host had disappeared<br />

at least 10 years previously, maintains a high and diverse inoculum of


12 DSE of Mediterranean mycorrhizal plants 215<br />

DSE fungi capable of associating with the ericaceous plant (Bergero et al.<br />

2003). DSE fungi have also been isolated from soil from northern temperate<br />

forests (Jumpponen and Trappe 1998a). However, it remains uncertain<br />

whether DSE fungi are endowed with “competitive saprophytic ability”<br />

(Garrett 1950, 1956), i.e. are actually efficient at saprotrophically decomposing<br />

organic debris in the complex soil environment, or whether they<br />

exist primarily as root associates of different plants, including ectomycorrhizal<br />

hosts (see Chap. 13 by Rice and Currah). Enzymatic activity may<br />

assist penetration into root tissues (Jumpponen and Trappe 1998a; Schulz<br />

et al. 2002).<br />

Whatever their actual free-living and saprotrophic abilities, when in association<br />

with host roots DSE isolates have been shown to increase host<br />

foliar P and N concentrations and plant biomass under some experimental<br />

conditions (Haselwandter and Read 1982; Fernando and Currah 1996;<br />

Jumpponen and Trappe 1998b; Jumpponen et al. 1998; Newsham 1999).<br />

These results suggest that at least some strains of DSE may have, from<br />

a functional point of view, a relationship with their host plants not dissimilar<br />

from that of conventional mycorrhizal symbionts. It should also be considered<br />

that variation in host response to classical mycorrhizal fungi may<br />

represent a continuum, ranging from parasitism to mutualism [Jumpponen<br />

2001; see Chaps. 16 (Brundrett) and 15 (Schulz)]. Involvement in nutrient<br />

and possibly water acquisition could be especially relevant in unfavourable<br />

environments exposed to droughts (Sengupta et al. 1989; Jumpponen et<br />

al. 1998). DSE are found extensively in xeric cold and warm environments<br />

(Read and Haselwandter 1981; Haselwandter and Read 1982; Kohn and<br />

Stasovski 1990; Barrow et al. 1997; Ruotsalainen et al. 2002), where they<br />

maybeevenmoreprevalentthanconventionalmycorrhizalfungi,colonising<br />

roots extensively with active structures (Barrow and Aaltonen 2001;<br />

Barrow and Osuna 2002; Barrow 2003). This suggests special adaptations<br />

to the harsh conditions of dry soil, possibly providing, when they occur<br />

concurrently with mycorrhizal symbionts, a back-up system during periods<br />

when the latter are inhibited by environmental conditions (Jumpponen<br />

and Trappe 1998a).<br />

Inoculation experiments in Earboreawith DSE isolates obtained from<br />

either this plant or Qilexhave shown that, irrespective of the host of origin,<br />

some strains are capable of forming typical intracellular hyphal coils,<br />

suggesting ericoid mycorrhizal behaviour (Bergero et al. 2000, 2003). Morphotype<br />

Sd9 isolates could also produce an unusual phenotype in the form<br />

of an additional hyphal net surrounding root epidermal cells (Fig. 12.2).<br />

While ecto- and ectoendo-mycorrhizal potential was reported for fungi of<br />

the DSE complex (Wilcox and Wang 1987a, 1987b; Ursic and Peterson 1997;<br />

Vrålstad et al. 2002b), the capacity to form endomycorrhizal structures<br />

such as coils had thus far not been described for the latter fungi. Such


216 M. Girlanda et al.<br />

Fig.12.2. Effects of Rhizopycnis vagum isolates of different origins on melon roots. Roots<br />

were rated for disease severity on the scale of Aegerter et al. (2000) [0 = no symptoms;<br />

1 = few lesions (covering < 10% of root), secondary root rot slight; 2 = rot of secondary<br />

roots or lesions covering approximately 25% of the root; 3 = lesions covering at least 50%<br />

of the root and dead secondary roots; 4 = general root rot, most of the root affected] after<br />

inoculation with R. vagum (5,000 cfu/g soil) at 25–28 ◦ C for 40–50 days. Monosporascus<br />

cannonballus and Acremonium cucurbitacearum, two other vine decline pathogens, were<br />

inoculated for comparison [30 cfu/g soil and 20,000 cfu/g soil, respectively (Aegerter et al.<br />

2000)]. Error bars indicate the standard deviation of the mean of ten observations in one<br />

growth chamber experiment. Different letters indicate differences significant at P


12 DSE of Mediterranean mycorrhizal plants 217<br />

gin. None of the tested isolates were able to form both kinds of mycorrhizal<br />

symbioses (Vrålstad et al. 2002b). Similarly, Sd2 and Sd9, the two morphotypesisolatedfromtheQ.<br />

ilex/ E. arborea host pair in Mediterranean plant<br />

communities, were only able to form typical ericoid mycorrhiza in the latterhostundertheconditionstested,despitethefactthatRAPDdatafor<br />

the Sd9 morphotype demonstrate the presence of the same genetic individual<br />

in both host plants (Bergero et al. 2000). Nonetheless, ectomycorrhizal<br />

behaviour on suitable hosts cannot be ruled out entirely, as suggested by<br />

high ITS sequence identity with an ectomycorrhizal symbiont of Kobresia<br />

myosuroides.<br />

Whatever their specific structural association with the root, in nature<br />

the multiple association potential of DSE fungi may favour inter-plant interactions,whichcaninturnaffectthediversityanddynamicsofplant<br />

communities. Mycelia of DSE fungi might interlink roots of different hosts,<br />

which could translocate nutrients via the hyphae of such shared fungal associates,<br />

similar to what has been shown in situations involving mycorrhizal<br />

fungi (Simard et al. 1997a, 1997b). Experiments with isotope tracers are<br />

obviously required to confirm such a hypothesis. A second possibility, not<br />

necessarily implying physical integrity and continuity of mycelia colonising<br />

different hosts, is that the ectomycorrhizal host provides a “reservoir” for<br />

mycorrhizal infection of other plants. Such a role as an efficient source of<br />

mycorrhizal inoculum for newly establishing seedlings would be especially<br />

relevant in highly disturbed habitats such as occur in Mediterranean environments.<br />

Results of isolation experiments from mature Q. ilex woodland<br />

soil have established survival of ericoid mycorrhizal fungi in the absence<br />

of the ericoid plant, which could facilitate secondary successions.<br />

Further experimentation is needed to verify the actual abilities of the<br />

Mediterranean DSE investigated thus far both in improving host growth<br />

and health and in forming typical mycorrhizal structures under natural<br />

conditions. In nature, these fungi face a variety of abiotic conditions as<br />

well as interactions with complex communities of microorganisms also<br />

interacting with the roots.<br />

A different behaviour was highlighted in the Halep pine/rosemary investigation<br />

(Girlanda et al. 2002). Identification of a DSE morphotype as<br />

Rhizopycnis vagum was unexpected since this fungus had thus far only<br />

been known as a root pathogen involved in cucurbit crop “vine declines”<br />

under agricultural conditions. R. vagum-specific PCR primers designed<br />

towards the ITS region have been developed to assist disease diagnosis<br />

(Ghignone et al. 2003). Association of R. vagum with unrelated hosts such<br />

as cucurbit crops and wild garrigue tree and shrub plants appears to differfromsituationsinwhichcroppathogensareendophyticinweedsin<br />

affected fields (Sinclair and Cerkauskas 1996). Pathogenic association has<br />

also been reported between Phialocephala fortinii and some host plants, as


218 M. Girlanda et al.<br />

revealed by resynthesis experiments under controlled conditions (Wilcox<br />

and Wang 1987b; Stoyke and Currah 1993; Fernando and Currah 1996). No<br />

disease, however, has so far been associated with DSE colonisation under<br />

natural conditions. Reports for R. vagum therefore raise several questions<br />

about the actual ecological plasticity of this fungus. Analyses of polymorphisms<br />

in single and multilocus genetic markers in pathogenic and<br />

endophytic isolates from cucurbit, P. halepensis and R. officinalis plants<br />

from Italy and Northern and Central America are currently underway to<br />

assess host-specific and geographical genetic variation within the fungus.<br />

Based on disease reaction in melon roots, the pathogenicity of endophytic<br />

R. vagum isolates from P. halepensis and R. officinalis wasconfirmedin<br />

greenhouses and growth chambers under different temperature regimes.<br />

Melon seedling inoculation demonstrated virulence on this host, establishing<br />

inherent pathogenic potential on a cucurbit plant, with at least some<br />

endophytic Italian isolates being more aggressive, under certain experimental<br />

conditions, than either of two pathogenic American isolates from<br />

diseased cucurbits (Fig. 12.2).<br />

Pathogenic behaviour has also been reported for Phomopsis / Diaporthe,<br />

another DSE morphotype from Halep pine and rosemary. Isolates of these<br />

genera are among the most common of the avirulent endophytes of both the<br />

above-ground organs and the roots of many plants, including trees, while<br />

others are known as plant pathogens (although mostly from epigeous plant<br />

organs) with a widespread occurrence (Sutton 1980; Alexopoulos et al.<br />

1996). Many fungal pathogens are often identified among the large numbers<br />

of fungi isolated from asymptomatic stems of woody hosts (Redlin and<br />

Carris 1996; Saikkonen et al. 1998). Disease symptoms were not apparent<br />

in the Halep pine and rosemary individuals sampled for DSE isolation. Endophytic<br />

fungi colonising asymptomatically plant parts may also include<br />

pathogens that have extended latency periods before development of disease,<br />

which occurs under conditions that induce host stress (Saikkonen et<br />

al. 1998, see Chap. 1 by Schulz and Boyle). Implicit in such a concept of<br />

latent pathogen is virulence, referring to the ability of the fungus to cause<br />

a disease in the particular hosts. Inoculation experiments in Petri dishes,<br />

containing mineral agar, of axenically cultured P. halepensis seedlings with<br />

DSE isolates identifiable as Phomopsis / Diaporthe did not result in disease<br />

symptoms, despite of experimental conditions not especially favourable<br />

to the plant, and the exposure to high fungal inoculum concentrations<br />

(unpublished data). Although the possibility of disease development in<br />

P. halepensis or R. officinalis under specific abiotic or biotic environmental<br />

conditions cannot be ruled out entirely, an alternative possibility is actual<br />

loss of virulence on these hosts and an inability to breach their defence barriers,<br />

while retaining some capacity of penetration into root tissues, which<br />

permits endophytic colonisation. Fungal endophytes of epigeous parts of


12 DSE of Mediterranean mycorrhizal plants 219<br />

both grasses and woody plants are closely related to pathogenic fungi, and<br />

are thought to have evolved from them via an extension of latency periods<br />

and a reduction of virulence (Schardl and Clay 1997; Saikkonen et al. 1998).<br />

DSE fungi could thus fit a definition of “saprotrophic symbionts”, living in<br />

close association with their hosts but confined by saprotrophy to the dead<br />

host tissues, to diffusates or even refractory structural components of such<br />

tissues, or to exudates from living host tissues, or even to organic material<br />

made available by other root associates (Cooke and Rayner 1984; Cooke<br />

and Whipps 1987).<br />

The results described here suggest that the outcome of the interaction<br />

between the same DSE fungus and different plant hosts, determining a conventional<br />

mycorrhizal, endophytic or even pathogenic association, may<br />

result from differences in fungal gene expression in response to the plant<br />

or differences in the ability of a plant to respond to the fungus, as has been<br />

suggested for other plant-fungus associations (Redman et al. 2001). This<br />

would fit well to a model of a finely tuned equilibrium between fungal virulence<br />

and plant defence characterising the fungal endophyte-plant host<br />

interaction, which could be described as a “balanced antagonism” (Schulz<br />

et al. 1999, 2002; see Chap. 1 by Schulz and Boyle).<br />

12.4<br />

Conclusions<br />

Molecular studies based on rDNA sequences are beginning to unravel<br />

the heterogeneity of the assemblage of DSE endophytes associated with<br />

different plants in different ecosystems. Analyses of sequence data from<br />

isolates from Mediterranean environments in Italy have widened the spectrum<br />

of DSE taxa known to associate with ecto- and endo-mycorrhizal<br />

hosts, pointing to broad taxonomic boundaries of these endophytes. Sequencing<br />

of rDNA coding regions has indicated affiliation to distant taxa<br />

within Ascomycota (distinct subclasses); however, taxonomic placement<br />

of these fungi has been achieved with varying degrees of resolution. ITSbased<br />

species-level identification of sterile fungi remains elusive in most<br />

cases, either because of matches with sequences from other unidentified<br />

fungi, or because of poor matches with all available sequences in the<br />

EMBL/GenBank/DDBJ databases. In spite of the daily increase in fungal<br />

sequences available in public databases, less than 1% of the estimated<br />

1.5 million fungal species are represented in public databases (Vilgalys<br />

2003), with large gaping holes for most fungal groups – Ascomycota included,<br />

with many families and even orders having no sequenced members<br />

to date. Misidentifications of named published sequences (upwards of 20%<br />

of the named sequences may be attributed to incorrectly named organisms;


220 M. Girlanda et al.<br />

Vilgalys 2003), and hence their unreliability, may represent another problem<br />

restricting feasibility of sequence-based identifications (see e.g. Bridge<br />

et al. 2003, 2004; Hawksworth 2004).<br />

Since intraspecific variation in the ITS region may attain ca. 15% in some<br />

fungal taxa (Egger and Sigler 1992; Seifert et al. 1995), conspecificity may<br />

be supposed with confidence only in cases where ITS sequence identity is<br />

quite high. For this reason, the amount of nucleotide divergence in single<br />

loci cannot in itself be used to define taxonomic rank, and phylogenetic<br />

species recognition relying on concordance between several independent<br />

gene genealogies (Taylor et al. 2000) may be a more valuable diagnostic<br />

tool. However, fungal non-ITS sequences of systematic value at the species<br />

level are scarce in databases.<br />

Regardless of their precise identity, however, database searches for DSE<br />

ITS sequences indicate that while some DSE fungi are possibly restricted<br />

to specific environments, in other cases the same fungus, or very closely<br />

related taxa, may occur in different biomes and unrelated hosts. The possibility<br />

that taxonomically diverse fungi have evolved not only different<br />

patterns of biome and host specificity, but also diverse functional significance<br />

for the plant is not unexpected. Data from inoculation experiments<br />

under semicontrolled conditions suggest that the range of DSE associates<br />

may vary from fungi with conventional mycorrhizal potential to fungi with<br />

pathogenic attributes. Findings for Mediterranean DSE isolates adds credence<br />

to the view of a possible continuum of mycorrhizal, endophytic or<br />

pathogenic root associations, depending on phylogenetic and life history<br />

constraints, geography, interactions with other species in the community<br />

and prevailing abiotic factors (Saikkonen et al. 1998; Brundrett 2002; see<br />

Chap. 16 by Brundrett). Assessing the ecological significance of information<br />

derived from axenic culture work with isolated DSE and the influences<br />

of plant and fungal genotypes, as well as local abiotic and biotic environments,<br />

on endophyte-host interactions under natural conditions is a major<br />

challenge for the future.<br />

Acknowledgements. Stefano Ghignone, Roberta Bergero, Cristina Mariani<br />

and Giacomo Tamietti contributed to the experimental work described<br />

in this chapter. Grants were provided by IPP-CNR Torino, UNITO 60%,<br />

CEBIOVEM.<br />

<strong>References</strong><br />

Addy HD, Hambleton S, Currah RS (2001) Distribution and molecular characterisation of<br />

the root endophyte Phialocephala fortinii along an environmental gradient in the boreal<br />

forest of Alberta. Mycol Res 104:1213–1221


12 DSE of Mediterranean mycorrhizal plants 221<br />

Aegerter BJ, Gordon TR, Davis RM (2000) Occurrence and pathogenicity of fungi associated<br />

with melon root rot and vine decline in California. Plant Dis 84:224–230<br />

Ahlich K, Sieber TN (1996) The profusion of dark septate endophytic fungi in nonectomycorrhizal<br />

fine roots of forest trees and shrubs. New Phytol 132:259–270<br />

Ahlich K, Rigling D, Holdenrieder O, Sieber TN (1998) Dark septate hyphomycetes in Swiss<br />

conifer forest soils surveyed using Norway-spruce seedlings as bait. Soil Biol Biochem<br />

30:1069–1075<br />

Alcorn JL, Irwin JAG (1987) Acrocalymma medicaginis gen. et sp. nov. causing root and<br />

crown rot of Medicago sativa in Australia. Trans Br Mycol Soc 88:163–167<br />

Alexopoulos CJ, Mims CW, Blackwell M (1996) Introductory mycology, 4th edn. Wiley,<br />

New York<br />

Allen MF (1991) The ecology of mycorrhizae. Cambridge University Press, Cambridge<br />

Allen TR, Millar T, Berch S, Berbee ML (2003) Culturing and direct DNA extraction find<br />

different fungi from the same ericoid mycorrhizal roots. New Phytol 160:255–272<br />

Armengol J, Pellicer I, Vicent A, Bruton BD, García-Jiménez J (2000) Rhizopycnis vagum<br />

DF Farr, un nuevo Coelomycete asociato a raíces de planta de melón con síntomas de<br />

colapso en España. Boletín de Sanidad Vegetal Plagas (1 ◦ Trimestre) 26:103–112<br />

Armengol J, Vicent A, Martínez-Culebras P, Bruton BD, García-Jiménez J (2003) Identification,<br />

occurrence and pathogenicity of Rhizopycnis vagum on muskmelon in Spain.<br />

Plant Pathol 52:68–73<br />

Bååth E, Söderström B (1980) Degradation of macromolecules by microfungi from different<br />

podzolic soil horizons. Can J Bot 58:422–425<br />

Barrow JR (2003) Atypical morphology of dark septate fungal root endophytes of Bouteloua<br />

in arid southwestern USA rangelands. Mycorrhiza 13:239–247<br />

Barrow JR, Aaltonen RE (2001) A method of evaluating internal colonisation of Atriplex<br />

canescens (Pursh) Nutt. roots by dark septate fungi and how they are influenced by<br />

physiological activity. Mycorrhiza 11:199–205<br />

Barrow JR, Osuna P (2002) Phosphorus solubilization and uptake by dark septate fungi in<br />

fourwing saltbush, Atriplex canescens (Pursh) Nutt. J Arid Environ 51:449–459<br />

Barrow JR, Havstad KM, McCaslin BD (1997) Fungal root endophytes in fourwing saltbush,<br />

Atriplex canescens, on arid rangelands of southwestern USA. Arid Soil Res Rehabil<br />

11:177–185<br />

Bergero R, Perotto S, Girlanda M, Vidano G, Luppi AM (2000) Ericoid mycorrhizal fungi<br />

are common root associates of a Mediterranean ectomycorrhizal plant (Quercus ilex).<br />

Mol Ecol 9:1639–1649<br />

Bergero R, Girlanda M, Bello F, Luppi AM, Perotto S (2003) Soil persistence and biodiversity<br />

of ericoid mycorrhizal fungi in the absence of the host plant in a Mediterranean<br />

ecosystem. Mycorrhiza 13:69–75<br />

Bridge PD, Roberts PJ, Spooner BM, Panchal G (2003) On the unreliability of published<br />

DNA sequences. New Phytol 160:43–48<br />

Bridge PD, Spooner BM, Roberts PJ (2004) Reliability and use of published sequence data.<br />

New Phytol 161:15–17<br />

Brundrett MC (2002) Tansley review no. 134. Coevolution of roots and mycorrhizas of land<br />

plants. New Phytol 154:275–304<br />

Bruton BD, Miller ME (1997a) Occurrence of vine decline diseases of muskmelon in<br />

Guatemala. Plant Dis 81:694<br />

Bruton BD, Miller ME (1997b) Occurrence of vine decline diseases of melons in Honduras.<br />

Plant Dis 81:696<br />

Bruton BD, Miller ME (1997c) The impact of vine declines on muskmelon production in<br />

Central America. Phytopathology 88:S120


222 M. Girlanda et al.<br />

Cairney JWG, Ashford AE (2002) Tansley review no. 135. Biology of mycorrhizal associations<br />

of epacrids (Ericaceae). New Phytol 154:275–304<br />

Caldwell BA, Jumpponen A, Trappe JM (2000) Utilization of major detrital substrates by<br />

dark-septate root endophytes. Mycologia 92:230–232<br />

Chambers SM, Liu G, Cairney JWG (2000) ITS rDNA sequence comparison of ericoid<br />

mycorrhizal endophytes from Woollsia pungens. Mycol Res 104:168–174<br />

Cooke RC, Rayner ADM (1984) Ecology of saprotrophic fungi. Longman, London<br />

Cooke RC, Whipps JM (1987) Saprotrophy, stress and symbiosis. In: Rayner ADM,<br />

Brasier CM, Moore D (eds) Evolutionary biology of the fungi. Cambridge University<br />

Press, Cambridge, pp 137–148<br />

Crayn DM, Quinn CJ (2000) The evolution of the atpβ-rbcL intergenic spacer in the epacrids<br />

(Ericales) and its systematic and evolutionary implications. Mol Phylogenet Evol<br />

16:238–252<br />

Currah RS, Tsuneda A (1993) Vegetative and reproductive morphology of Phialocephala<br />

fortinii (Hyphomycetes, Mycelium radicis atrovirens)inculture.TransMycolSocJapan<br />

34:345–356<br />

Currah RS, Sigler L, Hambleton S (1987) New records and new taxa of fungi from the<br />

mycorrhizae of terrestrial orchids of Alberta. Can J Bot 65:2473–2482<br />

Dahlberg A, Jonsson L, Nylund J-E(1997) Species diversity and distribution of biomass above<br />

and below ground among ectomycorrhizal fungi in an old-growth Norway spruce forest<br />

in south Sweden. Can J Bot 75:1323–1335<br />

Deacon JC (1973) Phialophora radicicola and Gaeumannomyces graminis on roots of grasses<br />

and cereals. Trans Br Mycol Soc 61:471–485<br />

Di Castri F, Mooney H (1973) Mediterranean type ecosystems. Springer, New York Berlin<br />

Heidelberg<br />

Egger KN, Sigler L (1992) Using molecular data to infer anamorph-teleomorph connections:<br />

the case of Scytalidium vaccinii and Hymenoscyphus ericae. In: Reynolds DR, Taylor JW<br />

(eds) The fungal holomorph: mitotic, meiotic and pleomorphic speciation in fungal<br />

systematics. CAB International, Wallingford, pp 141–146<br />

Farr DF, Miller ME, Bruton BD (1998) Rhizopycnis vagum genetspnov,anewcoelomycetous<br />

fungus from roots of melons and sugarcane. Mycologia 90:290–296<br />

Fernando AA, Currah RS (1995) Leptodontidium orchidicola (Mycelium Radicis Atrovirens<br />

complex): aspects of its conidiogenesis and ecology. Mycotaxon 54:287–294<br />

Fernando AA, Currah RS (1996) A comparative study of the effects of the root endophytes<br />

Leptodontidium orchidicola and Phialocephala fortinii (Fungi Imperfecti) on the growth<br />

of some subalpine plants in culture. Can J Bot 74:1071–1078<br />

Gams W (1963) Mycelium radicis atrovirens in forest soils, isolation from soil microhabitats<br />

and identification. In: Doeksen J, van der Drift J (eds) Soil organisms. North-Holland<br />

Publications, Amsterdam, pp 176–182<br />

Garrett SD (1950) Ecology of the root-inhabiting fungi. Biological Reviews 25:220–254<br />

Garrett SD (1956) Biology of root-infecting fungi. Cambridge University Press, Cambridge<br />

Ghignone S, Tamietti G, Girlanda M (2003) Development of specific PCR primers for identification<br />

and detection of Rhizopycnis vagum. Eur J Plant Pathol 109:861–870<br />

Girlanda M, Luppi-Mosca AM (1995) Microfungi associated with ectomycorrhizae of Pinus<br />

halepensis Mill. Allionia 33:93–98<br />

Girlanda M, Ghignone S, Luppi AM (2002) Diversity of root-associated fungi of two Mediterranean<br />

plants. New Phytol 155:481–498<br />

Grünig CR, Sieber TN, Holdenrieder O (2001) Characterisation of dark septate endophytic<br />

fungi (DSE) using inter-simple-sequence-repeat-anchored polymerase chain reaction<br />

(ISSR-PCR) amplification. Mycol Res 105:24–32


12 DSE of Mediterranean mycorrhizal plants 223<br />

Grünig CR, Sieber TN, Rogers SO, Holdenrieder O (2002a) Genetic variability among strains<br />

of Phialocephala fortinii and phylogenetic analysis of the genus Phialocephala based on<br />

rDNA ITS sequence comparisons. Can J Bot 80:1239–1249<br />

Grünig CR, Sieber TN, Rogers SO, Holdenrieder O (2002b) Spatial distribution of dark<br />

septate endophytes in a confined forest plot. Mycol Res 106:832–840<br />

Gwinne BJ, Davis RM, Gordon TR (1997) Occurrence and pathogenicity of fungi associated<br />

with melon vine decline in California. Phytopathology 87:S37<br />

Hambleton S, Currah RS (1997) Fungal endophytes from the roots of alpine and boreal<br />

Ericaceae. Can J Bot 75:1570–1581<br />

Harney SK, Rogers SO, Wang CJK (1997) Molecular characterisation of dematiaceous root<br />

endophytes. Mycol Res 101:1397–1404<br />

Harrington TS, McNew DL (2003) Phylogenetic analysis places the Phialophora-like<br />

anamorph genus Cadophora in the Helotiales. Mycotaxon 87:141–152<br />

Haselwandter K (1983) Pectic enzymes produced by fungal root associates of alpine plants.<br />

Phyton 28:55–64<br />

Haselwandter K, Read DJ (1982) The significance of root-fungus associations in two Carex<br />

species of high-alpine plant communities. Oecologia 53:352–354<br />

Hawksworth DL (2004) ‘Misidentifications’ in fungal DNA sequence databanks. New Phytol<br />

161:13–15<br />

Holdenrieder O, Sieber TN (1992) Fungal associations of serially washed, healthy, nonmycorrhizal<br />

roots of Picea abies. Mycol Res 96:151–156<br />

Jumpponen A (1999) Spatial distribution of discrete RAPD phenotypes of a root endophytic<br />

fungus, Phialocephala fortinii, at a primary successional site on a glacier forefront. New<br />

Phytol 141:333–344<br />

Jumpponen A (2001) Dark septate endophytes – are they mycorrhizal? Mycorrhiza<br />

11:207–211<br />

Jumpponen A, Trappe JM (1998a) Dark septate endophytes: a review of facultative biotropic<br />

root-colonising fungi. New Phytol 140:295–310<br />

Jumpponen A, Trappe JM (1998b) Performance of Pinus contorta inoculated with two<br />

strains of root endophytic fungus Phialocephala fortinii: effects of resynthesis system<br />

and glucose concentration. Can J Bot 76:1205–1213<br />

Jumpponen A, Mattson KG, Trappe JM (1998) Mycorrhizal functioning of Phialocephala<br />

fortinii: interactions with soil nitrogen and organic matter. Mycorrhiza 7:261–265<br />

Kirk PM, Cannon PF, David JC, Stalpers JA (2001) Ainsworth and Bisby’s dictionary of the<br />

fungi, 9th edn. CAB International, Wallingford<br />

Kohn L, Stasovski E (1990) The mycorrhizal status of plants at Alexandra Fiord, Ellesmere<br />

Island, Canada, a high Arctic site. Mycologia 82:23–35<br />

Kovacs GM, Szigetvari C (2002) Mycorrhizae and other root-associated fungal structures of<br />

the plants of a sandy grassland on the Great Hungarian Plain. Phyton-Annu Rev Bot A<br />

42:211–223<br />

LoBuglio KF, Berbee ML, Taylor JW (1996) Phylogenetic origins of the asexual mycorrhizal<br />

symbiont Cenococcum geophilum Fr. and other mycorrhizal fungi among the<br />

ascomycetes. Mol Phylogenet Evol 6:287–294<br />

McLean CB, Cunnington JH, Lawrie AC (1999) Molecular diversity within and between<br />

ericoid endophytes from the Ericaceae and Epacridaceae. New Phytol 144:351–358<br />

Melin E (1922) On the mycorrhizas of Pinus sylvestris L. and Picea abies Karst.Apreliminary<br />

note. J Ecol 9:254–257<br />

Melin E (1923) Experimentelle Untersuchungen über die Konstitution und Okologie der<br />

Mykorrhizen von Pinus silvestris und Picea abies. Mykologische Untersuchungen und<br />

Berichte von R. Falck 2:73–331


224 M. Girlanda et al.<br />

Merteley JC, Martyn RD, Miller ME, Bruton BD (1991) Role of Monosporascus cannonballus<br />

and other fungi in a root rot/vine decline of muskmelon. Plant Dis 75:1133–1137<br />

Miller ME, Bruton BD, Farr DF (1996) Association of a Stagonospora-like fungus on roots<br />

of melons exhibiting vine decline symptoms. Phytopathology 86:S3<br />

Monreal M, Berch SM, Berbee M (1999) Molecular diversity of ericoid mycorrhizal fungi.<br />

Can J Bot 77:1580–1594<br />

Montuschi C (2001) Collasso delle cucurbitacee: una malattia da indagare. Agricoltura<br />

30:31–33<br />

Newsham KK (1999) Phialophora graminicola, a dark septate fungus, is a beneficial associate<br />

of the grass Vulpia ciliata ssp. ambigua. New Phytol 144:517–524<br />

O’Dell TE, Massicotte HB, Trappe JM (1993) Root colonisation of Lupinus latifolius Agardh.<br />

and Pinus contorta Dougl. by Phialocephala fortinii Wang&Wilcox.NewPhytol<br />

124:93–100<br />

Peyronel B (1924) Prime ricerche sulla micorize endotrofiche e sulla microflora radicola<br />

normale delle fanerogame. Rivista di Biologia 6:17–53<br />

Pignatti E, Pignatti S, Ladd PG (2002) Comparison of ecosystems in the Mediterranean<br />

Basin and Western Australia. Plant Ecol 163:177–186<br />

Porta-Puglia A, Pucci N, Di Gianbattista G, Infantino A (2001) First report of Rhizopycnis<br />

vagum associated with tomato roots in Italy. Plant Dis 85:1210<br />

Read DJ (1991) Experimental simplicity versus natural complexity in mycorrhizal systems.<br />

In: Fontana A (ed) Funghi, piante e suolo. Consiglio Nazionale delle Ricerche, Turin,<br />

pp 75–104<br />

Read DJ (2000) Links between genetic and functional diversity – a bridge too far? New<br />

Phytol 145:363–365<br />

Read DJ, Haselwandter K (1981) Observations on the mycorrhizal status of some alpine<br />

plant communities. New Phytol 88:341–352<br />

Redlin SC, Carris LM (1996) Endophytic fungi in grasses and woody plants. APS Press,<br />

St Paul, MN<br />

Redman RS, Dunigan DD, Rodriguez RJ (2001) Fungal symbiosis from mutualism to parasitism:<br />

who controls the outcome, host or invader? New Phytol 151:705–7161<br />

Richard C, Fortin JA (1973) The identification of Mycelium radicis atrovirens. CanJBot<br />

51:2247–2248<br />

Ruotsalainen AL, Vare H, Vestberg M (2002) Seasonality of root fungal colonisation in<br />

low-alpine herbs. Mycorrhiza 12:29–36<br />

Saikkonen K, Faeth SH, Helander M, Sullivan TJ (1998) Fungal endophytes: a continuum of<br />

interactions with host plants. Annu Rev Ecol Syst 29:319–343<br />

Schadt CW, Mullen RB, Schmidt SK (2001) Isolation and phylogenetic identification of<br />

a dark-septate fungus associated with the alpine plant Ranunculus adoneus.NewPhytol<br />

150:747–755<br />

Schardl CL, Clay K (1997) Evolution of mutualistic endophytes from plant pathogens. In:<br />

Carroll GC, Tudzynski P (eds) The Mycota V: Plant relationships, part B. Springer, Berlin<br />

Heidelberg New York, pp 221–238<br />

Schild DE, Kennedy A, Stuart MR (1988) Isolation of symbiont and associated fungi from<br />

ectomycorrhizas of Sitka spruce. Eur J Forest Pathol 18:51–61<br />

Schulz B, Römmert AK, Dammann U, Aust HJ, Strack D (1999) The endophyte-host interaction:<br />

a balanced antagonism? Mycol Res 103:1275–1283<br />

Schulz B, Boyle C, Draeger S, Römmert AK, Krohn K (2002) Endophytic fungi: a source of<br />

novel biologically active secondary metabolites. Mycol Res 106:996–1004<br />

Seifert KA, Wingfield BD, Wingfield MJ (1995) A critique of DNA sequence analysis in<br />

the taxonomy of filamentous Ascomycetes and ascomycetous anamorphs. Can J Bot<br />

73:S760–S767


12 DSE of Mediterranean mycorrhizal plants 225<br />

Sengupta A, Chakraborty DC, Chaudhuri S (1989) Do septate endophytes also have a mycorrhizal<br />

function for plants under stress? In: Mahadevan A, Raman N, Natarajan K (eds)<br />

Mycorrhizae for Green Asia. Proceedings of the 1st Asian Conference on Mycorrhizae,<br />

Nadras, India, pp 169–174<br />

Sharples JM, Chambers SM, Meharg AA, Cairney JWG (2000) Genetic diversity of rootassociated<br />

fungal endophytes from Calluna vulgaris at contrasting field sites. New<br />

Phytol 148:153–162<br />

Shoemaker RA, Babcock CE, Irwin JAG (1990) Massarina walkeri n. sp., the teleomorph of<br />

Acrocalymma medicaginis from Medicago sativa contrasted with Leptosphaeria pratensis,<br />

L. weimerin.sp.&L. viridella. Can J Bot 69:569–573<br />

Simard SW, Jones MD, Durall DM, Perry DA, Myrold DD, Molina R (1997a) Reciprocal<br />

transfer of carbon isotopes between ectomycorrhizal Betula papyrifera and Pseudotsuga<br />

menziesii New Phytol 137:529–542<br />

Simard SW, Perry DA, Jones MD, Myrold DD, Durall DM, Molina R (1997b) Net transfer of<br />

carbon between ectomycorrhizal tree species in the field. Nature 388:579–582<br />

Sinclair JB, Cerkauskas RF (1996) Latent infection vs endophytic colonisation by fungi. In:<br />

Redlin SC, Carris LM (eds) Endophytic fungi in grasses and woody plants. APS Press,<br />

St Paul, MN, pp 3–29<br />

Stoyke G, Currah RS (1991) Endophytic fungi from the mycorrhizae of alpine ericoid plants.<br />

Can J Bot 69:347–352<br />

Stoyke G, Currah RS (1993) Resynthesis in pure culture of a common sub-alpine fungus-root<br />

association using Phialocephala fortinii and Menziesia ferruginea (Ericaceae). Arctic<br />

Alpine Res 25:189–193<br />

Stoyke G, Egger KN, Currah RS (1992) Characterisation of sterile endophytic fungi from<br />

the mycorrhizae of subalpine plants. Can J Bot 70:2009–2016<br />

Summerbell RC (1989) Microfungi associated with the mycorrhizal mantle and adjacent<br />

microhabitats within the rhizosphere of black spruce. Can J Bot 67:1085–1095<br />

Sutton BC (1980) The Coelomycetes – Fungi Imperfecti with pycnidia, acervuli and stromata.<br />

Commonwealth Mycological Institute, Kew<br />

Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS, Fisher MC (2000)<br />

Phylogenetic species recognition and species concepts in fungi. Fungal Genet Biol<br />

31:21–32<br />

Taylor DL, Bruns TD, Leake JR, Read DJ (2002) Mycorrhizal specificity and function in mycoheterotrophic<br />

plants. In: Van der Heijden MGA, Sanders I (eds) Mycorrhizal ecology.<br />

Springer, Berlin Heidelberg New York, pp 375–413<br />

Ursic M, Peterson RL (1997) Morphological and anatomical characterisation of ectomycorrhizas<br />

and ectendomycorrhizas on Pinus strobus seedlings in a southern Ontario<br />

nursery. Can J Bot 75:2057–2072<br />

Vandenkoornhuyse P, Baldauf SL, Leyval C, Straczek J, Young JPW (2002) Extensive fungal<br />

diversity in plant roots. Science 295:2051<br />

Vilgalys R (2003) Taxonomic misidentification in public DNA databases. New Phytol 160:4–5<br />

Villarreal-Ruiz L, Anderson IC, Alexander IJ (2004) Interaction between an isolate from<br />

the Hymenoscyphus ericae aggregate and roots of Pinus and Vaccinium.NewPhytol<br />

164:183–192<br />

Vrålstad T (2004) Are ericoid and ectomycorrhizal fungi part of a common guild? New<br />

Phytol 164:7–10<br />

Vrålstad T, Fossheim T, Schumacher T (2000) Piceirhiza bicolorata – the ectomycorrhizal<br />

expression of the Hymenoscyphus ericae aggregate? New Phytol 145:549–563<br />

Vrålstad T, Myhre E, Schumacher T (2002a) Molecular diversity and phylogenetic affinities<br />

of symbiotic root-associated ascomycetes of the Helotiales in burnt and metal polluted<br />

habitats. New Phytol 155:131–148


226 M. Girlanda et al.<br />

Vrålstad T, Schumacher T, Taylor AFS (2002b) Mycorrhizal synthesis between fungal strains<br />

of the Hymenoscyphus ericae aggregate and potential ectomycorrhizal and ericoid hosts.<br />

New Phytol 153:143–152<br />

Wang CJK, Wilcox HE (1985) New species of ectendomycorrhizal and pseudomycorrhizal<br />

fungi: Phialophora finlandia, Chloridium paucisporum,andPhialocephala fortinii.Mycologia<br />

77:951–958<br />

Wilcox HE, Wang CJK (1987a) Ectomycorrhizal and ectendomycorrhizal associations of<br />

Phialophora finlandia with Pinus resinosa, Picea rubens, and Betula alleghaniensis. Can<br />

J For Res 17:976–990<br />

Wilcox HE, Wang CJK (1987b) Mycorrhizal and pathological associations of dematiaceous<br />

fungi in roots of 7-month-old tree seedlings. Can J For Res 17:884–899<br />

Yu TEJ-C, Egger KN, Peterson RL (2001) Ectendomycorrhizal associations – characteristics<br />

and functions. Mycorrhiza 11:167–177


13<br />

13.1<br />

Introduction<br />

Oidiodendron maius:<br />

Saprobe in Sphagnum Peat,<br />

Mutualist in Ericaceous Roots?<br />

Adrianne V. Rice, Randolph S. Currah<br />

Oidiodendron maius Barron is a hyphomycete species isolated from peat,<br />

soil, decaying organic matter, and plant roots throughout temperate ecosystems,<br />

including peatlands, forests, and heathlands (e.g. Barron 1962; Nordgren<br />

et al. 1985; Schild et al. 1988; Nilsson et al. 1992; Hambleton et al. 1998;<br />

Qian et al. 1998; Lumley et al. 2001; Thormann 2001; Thormann et al. 2001,<br />

2004; Tsuneda et al. 2001; Rice and Currah 2002; Rice et al. 2006). In pure<br />

culture, colonies are white due to the presence of abundant arthroconidia<br />

that develop in chains at the apex of thick-walled, melanized erect conidiophores<br />

30–500 µm tall (Fig. 13.1a). Conidia are thin-walled, subglobose to<br />

elongate or irregular, 2-5x1–2.5 µm, and have an asperulate perispore (Rice<br />

and Currah 2001) (Fig. 13.1b).<br />

A sexual state is unknown but morphological characters indicate a close<br />

affiliation to other taxa in the Myxotrichaceae (a cleistothecial family in<br />

the Helotiales; Tsuneda and Currah 2004): six teleomorph species within<br />

the Myxotrichaceae have Oidiodendron states (Hambleton et al. 1998; Rice<br />

and Currah 2005) and sterile ascomata with peridial elements resembling<br />

those formed by species of Myxotrichum canbeinducedwhenthespeciesis<br />

grown on autoclaved lichen (Rice and Currah 2002) (Fig. 13.1c). Molecular<br />

evidence confirms the position of O. maius among other species of Oidiodendron<br />

and their teleomorphic counterpart, Myxotrichum (Hambleton et<br />

al. 1998).<br />

The distribution of O. maius seems to parallel that of members of the Ericaceae<br />

(blueberries, cranberries, rhododendrons, etc.), a family that often<br />

dominates the vegetation in arctic and alpine meadows, temperate heathlands,<br />

the understory in boreal forests and peatlands (Hambleton 1998;<br />

Chambers et al. 2000; Hambleton and Currah 2000), and Mediterranean<br />

ecosystems (Perotto et al. 1995). This shared distribution pattern may be<br />

Adrianne V. Rice: Northern Forestry Centre, Canadian Forest Service, Natural Resources<br />

Canada, 5320-122 St., Edmonton, AB, T6H 3S5 Canada, E-mail: ARice@NRCan.gc.ca<br />

Randolph S. Currah: Department of Biological Sciences, University of Alberta, Edmonton,<br />

AB, T6G 2E9, Canada<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


228 A.V. Rice, R.S. Currah<br />

Fig.13.1. a–c Morphology of Oidiodendron maius in axenic culture. a Tall, erect, melanized<br />

conidiophores and small, hyaline arthroconidia of O. maius (UAMH 9749) viewed under<br />

light microscopy. b Chains of arthroconidia of O. maius (UAMH 8920), showing asperulate<br />

ornamentation visible under scanning electron microscopy. c Sterile peridial elements<br />

produced by O. maius (UAMH 9749) on thalli of Cladonia mitis. The cage-like peridial<br />

elements resemble the cleistothecia of species of Myxotrichum. Bars a 40 µm, b 5µm,c 80 µm<br />

based on a predilection in both plants and fungus for acidic, nutrientpoor,<br />

organic soils; it is also possible that O. maius has some degree of<br />

dependency on a mycorrhizal or at least root-endophytic relationship with<br />

ericaceous plants.<br />

Thus, two hypotheses can be advanced to explain the distribution of<br />

O. maius. The first hypothesis, that it is a competitive and effective saprobe<br />

on acidic, organic soils, is supported by its optimal growth in culture on<br />

acidic growth media (Rice and Currah 2001, 2005), its ability to grow and<br />

sporulate readily on Sphagnum plants (Rice and Currah 2002), and its<br />

ability to produce enzymes that degrade the cell walls of Sphagnum leaves<br />

in vitro (Tsuneda et al. 2001; Rice et al. 2006). The second hypothesis, that<br />

O. maius is a mycorrhizal endophyte of ericaceous shrubs, is supported by<br />

its frequent isolation from ericaceous roots (e.g. Douglas et al. 1989; Perotto<br />

et al. 1995, 1996; Hambleton and Currah 1997; Currah et al. 1999; Monreal


13 Saprobe and mutualist 229<br />

et al. 1999; Chambers et al. 2000; Johansson 2001; Usuki et al. 2003) and<br />

by observations that it can form typical ericoid mycorrhizal infection units<br />

when reinoculated on Ericaceae grown in culture (e.g. Douglas et al. 1989;<br />

Xiao and Berch 1995, 1999; Monreal et al. 1999).<br />

Since Douglas et al. (1989) described the ericoid mycorrhizas formed<br />

by O. maius in Rhododendron, most reports of O. maius have been from<br />

ericaceous plants. Its role in these associations has been considered one in<br />

which the plant derives some nutritional benefit (e.g. Perotto et al. 1995,<br />

1996; Hambleton and Currah 1997, 2000; Johanson 2001). Oddly, the benefits<br />

accruing to the fungus in these relationships are rarely considered,<br />

possibly because they are considered secondary to the needs of the plant<br />

but perhaps also because they are difficult to determine (Douglas and Smith<br />

1989). Unlike arbuscular mycorrhizal fungi and many nutritionally fastidious<br />

ectomycorrhizal basidiomycetes that are difficult to grow in culture<br />

or which require the addition of complex compounds in artificial media,<br />

O. maius does not display stringent demands for host-derived sugars and<br />

or complex growth factors, and grows readily in culture on many types of<br />

natural materials, including lichen thalli, Sphagnum gametophytes, context<br />

tissues of polypores, and on artificial growth media. In the absence of<br />

nutritional dependency, the benefits to the fungus in a mycorrhizal or root<br />

endophytic relationship are usually speculative and a number of possibilities<br />

could be considered. For example, the relationship may provide the<br />

fungus with a carbon source, growth factors, habitat, a preemptive position<br />

as a consumer of senescent tissue, or a competitive advantage over other<br />

saprobes. Alternatively, the fungus may not benefit from the association,<br />

with host plants exploiting their fungal partners, as in orchid mycorrhizas<br />

(Rasmussen 1995). In summary, there are two possible explanations for<br />

the distribution of O. maius: i.e. it may be a saprobe adapted to acidic<br />

conditions and enzymatically equipped to digest the intractable materials<br />

that accumulate in these areas, or it may be a type of root endophyte, possibly<br />

a mycorrhizal one, that requires its host plants to thrive in its habitats.<br />

These two suggested ecological roles are not necessarily mutually exclusive,<br />

i.e. the species could occupy both mycorrhizal and saprobic niches within<br />

suitable ecosystems.<br />

OnetypeofecosystemthatmaybesupportingO. maius both as a saprobe<br />

and a mycorrhizal associate is acidic Sphagnum peatlands, found throughout<br />

the circumboreal region. These peatlands include bogs and fens with<br />

an understory of ericaceous shrubs, including Rhododendron, Andromeda,<br />

and Vaccinium species, and a thick ground layer of Sphagnum spp. (e.g.<br />

Svensson 1995; Vitt et al. 1996; Hoosbeek et al. 2001). In Canada, many peatlands<br />

have a canopy of coniferous trees rooted in the Sphagnum (Vitt et al.<br />

1996; Piercey et al. 2002). Bogs have a dense canopy of black spruce (Picea<br />

mariana) while poor fens are dominated by black spruce and larch (Larix


230 A.V. Rice, R.S. Currah<br />

spp.) (Vitt et al. 1996). European peatlands tend to have open canopies with<br />

few trees but, as in Canada, the common tree species in European peatlands<br />

are spruce (Picea abies) and larch (e.g. Peteet et al. 1998). Ericoid mycorrhizal<br />

fungi, including O. maius, have been isolated from ectomycorrhizas<br />

of conifers growing in such peatlands (Summerbell 1987; Schild et al. 1988;<br />

Perotto et al. 1995; Qian et al. 1998; Vrålstad et al. 2000); in some cases,<br />

O. maius was the most abundant sporulating species isolated from the roots<br />

(e.g. roots of sitka spruce sampled from blanket bogs; Schild et al. 1988). It<br />

has been proposed that these fungi may form associations with the roots<br />

ofconifersandericaceousshrubsinpeatlands,anddegradetheSphagnum<br />

matrix (Piercey et al. 2002).<br />

In this chapter, we first review the evidence suggesting that O. maius is<br />

a saprobe and then the evidence that it is an ericoid mycorrhizal fungus.<br />

Finally, we discuss why isolation records of this Helotialean anamorph point<br />

towards its simultaneous occupation of two apparently distinct niches.<br />

13.2<br />

Oidiodendron maius as a Saprobe<br />

From 1962 to 1989, Oidiodendron maius was known only from scattered<br />

records from soils and other decaying organic debris (Barron 1962; Nordgren<br />

et al. 1985), where it was presumed to occupy a saprobic niche, and<br />

from the ectomycorrhizal root tips of sitka spruce (Schild et al. 1988). Although<br />

Schild et al. (1988) considered O. maius a cortical parasite of spruce,<br />

evidence of parasitism was not presented; instead, the evidence suggested<br />

that O. maius inhibited root pathogens, including Phytophthora cinnamomi<br />

and Heterobasidium annosum. The inhibitory activity of O. maius towards<br />

root pathogens was also suggested by Qian et al. (1998), who found that<br />

O. maius was a dominant inhabitant of the ectomycorrhizal root tips of<br />

Norway spruce under acidified conditions.<br />

ThefirstreportofO. maius from the roots of ericaceous shrubs appeared<br />

in 1989, when it was isolated from ericoid mycorrhizal roots of Rhododendron<br />

(Douglas et al. 1989). Since then most records of O. maius have been<br />

based on isolates from presumably healthy ericaceous roots (e.g. Hambleton<br />

and Currah 1997; Currah et al. 1999; Monreal et al. 1999; Chambers et<br />

al. 2000; Usuki et al. 2003), but records from other substrates continue to<br />

appear (Nilsson et al. 1992; Qian et al. 1998; Lumley et al. 2001; Thormann<br />

et al. 2001, 2004; Rice and Currah 2002; Rice et al. 2006). The fungus is relatively<br />

easy to isolate from roots because it grows rapidly on artificial media,<br />

showing increases in colony radius of up to 1 mm/day on cornmeal agar<br />

(CMA) at pH 3 during periods of maximal growth (Rice and Currah 2001).<br />

The fungus has been shown to degrade a variety of carbon and nitrogen


13 Saprobe and mutualist 231<br />

sources including tannic acid (a soluble phenolic compound), cellulose,<br />

starch (Rice and Currah 2001, 2005; Thormann et al. 2002), chitin, pectin<br />

(Rice and Currah 2001, 2005), and TWEEN 20, a lipid-based detergent (Rice<br />

and Currah 2005).<br />

The presence of chitinases suggests that O. maius may obtain nutrients<br />

(e.g. nitrogen) from polymeric glucosamines found in insects and fungi.<br />

Oidiodendron maius grows luxuriantly on lichen (Rice and Currah 2002)<br />

and on the context tissue of the basidiocarps (polypores) of larger wood<br />

decay fungi (e.g. Fomitopsis pinicola). Oidiodendron maius also grows and<br />

sporulates abundantly on a lipid-rich growth medium with TWEEN 20.<br />

There are no data suggesting that O. maius is a fungal or arthropod parasite,<br />

but it could be a saprobe on materials rich in lipids and chitin, such as the<br />

remains of fungi and microfauna.<br />

This suite of cultural characteristics is more indicative of a saprobic<br />

lifestyle rather than a mycorrhizal one that relies on biotrophically derived<br />

photosynthates (Hutchison 1990, 1991). Alternatively, because ecto- and<br />

ericoid-mycorrhizal fungi have also been shown to degrade a variety of<br />

organic substrates (e.g. Bajwa and Read 1985; Northup et al. 1995; Bending<br />

and Read 1996, 1997; Aerts 2002; Leake et al. 2002; Olsson et al. 2002;<br />

Simard et al. 2002), these abilities may enable the absorption and transfer<br />

of organic or non-mineralised nutrients directly from the substrate to the<br />

cytoplasm of a host plant, effectively short-circuiting the mineralisation<br />

steps in nutrient cycling (‘organicization’) (Northup et al. 1995; Aerts 2002;<br />

Leake et al. 2002). In this instance, it is the host that derives benefit, and<br />

reciprocity for the fungus is not evident.<br />

The ability to sporulate in pure culture is much less common for mycorrhizal<br />

fungi than saprobes, with all arbuscular and most ectomycorrhizal<br />

fungi unable to reproduce in the absence of their hosts (Read 2002). Endorhizal<br />

fungi allied to the Helotiales are notable exceptions (Addy et<br />

al. 2005). For example, Rhizoscyphus ericae, an ericoid mycorrhizal fungus<br />

that also forms ectomycorrhizas (Vrålstad et al. 2000, 2002), produces<br />

chains of arthroconidia in culture, and in rare instances has formed discocarps<br />

(Hambleton et al. 1999). Unlike O. maius, R. ericae is unknown<br />

from non-mycorrhizal sources, although this may be due to its variable<br />

cultural morphology, and the concomitant difficulties in making definitive<br />

identification of this fungus (Hambleton and Currah 1997; Hambleton and<br />

Sigler 2005).<br />

As with other microfungi, assumptions about distribution and habitat<br />

preferences are biased by isolation protocols, expertise in identification,<br />

and by the nature and objective of the reports in which taxa are listed.<br />

Nevertheless, peat is a likely substrate on which to find O. maius because<br />

it is acidic and rich in many of the organic compounds, including tannic<br />

acid, cellulose, pectin, and chitin, that O. maius is able to degrade. The


232 A.V. Rice, R.S. Currah<br />

first record of O. maius was from “peat soil” (Barron 1962) and subsequent<br />

reports of this species from peat (e.g. Nilsson et al. 1992; Thormann et al.<br />

2001, 2004; Rice and Currah 2002, Rice et al. 2006) remain more common<br />

than reports from other organic debris, such as wood (Lumley et al. 2001).<br />

O. maius was more abundant in ectomycorrhizal root tips of spruce in<br />

blanket bogs than in mineral woodland soils (Schild et al. 1988) and in<br />

acidified rather than limed soils (Qian et al. 1998), further supporting an<br />

apparent preference for acidic substrates.<br />

The scant isolation data from other materials is possibly the result of<br />

the biases mentioned above. For example, when Rice and Currah (2002)<br />

compared the isolation frequency of this taxon using agar media and moist<br />

chambering, O. maius was the most abundant sporulating species appearing<br />

directly on Sphagnum peat,butitwasonlyrarelyencounteredwhenthe<br />

same peat was placed on agar media (Rice and Currah 2002). O. maius was<br />

not isolated from ectomycorrhizal root tips when benomyl was added to<br />

the isolation medium (Schild et al. 1988). O. maius is capable of growing on<br />

benomyl-amended media (Rice and Currah 2002) but may not have been<br />

able to compete with basidiomycetes and other fungi favoured by the addition<br />

of this selective antifungal agent. Growth and sporulation of O. maius<br />

is restricted on rich artificial media, including the potato dextrose and<br />

malt extract agars that are commonly used to isolate fungi from substrates,<br />

further biasing against its recovery.<br />

The isolation history of O. maius coupled with in vitro studies of its<br />

behaviour on Sphagnum peat strongly suggest that O. maius is abundant<br />

in this material and may degrade large quantities of the substrate under<br />

natural conditions (Thormann 2001; Piercey et al. 2002; Rice and Currah<br />

2002, Rice et al. 2006). Three studies have shown that O. maius can cause<br />

significant mass losses of Sphagnum in vitro, ranging from 2–3% (Thormann<br />

2001) to 10–12% (Piercey et al. 2002) and from negligible to almost<br />

50% (Rice et al. 2006). These mass loss data may differ because intact Sphagnum<br />

was used by Thormann (2001) and ground Sphagnum by Piercey et<br />

al. (2002). Differences may also be due to the strains of O. maius used,<br />

as suggested by the wide range observed by Rice et al. (2006) for three<br />

different strains. Piercey et al. (2002) compared mass losses caused by two<br />

isolates of O. maius (UAMH 8919, 8920) with other ericoid mycorrhizal<br />

fungi (R. ericae and a non sporulating white-to-grey fungus designated<br />

“VWT”) from the roots of peatland and heathland Ericaceae and found<br />

that the strains of O. maius caused the greatest losses. Thormann (2001)<br />

found that the mass losses caused by O. maius (UAMH 9749) were intermediate<br />

among five saprobic hyphomycetes [O. maius, Acremonium cf.<br />

curvulum (identified later as Pochonia bulbillosa, Thormann et al. 2004),<br />

Penicillium thomii, O. scytaloides (= O. chlamydosporicum sensu, Rice and<br />

Currah 2005), and Trichoderma viride] and greater than an unidentified


13 Saprobe and mutualist 233<br />

basidiomycete. Tsuneda et al. (2001) used scanning electron microscopy<br />

to compare the ultrastructural patterns of decay of Sphagnum caused by<br />

O. maius and P. bulbillosa. Sphagnum cell walls are analogous to wood,<br />

Fig.13.2. a–e Degradation of Sphagnum fuscum leaf cell walls by Oidiodendron maius<br />

(UAMH 9749; Tsuneda et al. 2001). a Affected cell wall showing finely wavy deformations<br />

(arrows). b Severely distorted leaf cell wall (arrow). Note autolysing hypha (arrowhead)and<br />

degraded leaf cell wall in the immediate vicinity. H Hypha, C conidia. c Localized voids<br />

(arrows)andhyphae(H) emerging through the leaf cell wall. d More or less simultaneous<br />

degradation of the leaf cell wall by a hypha (H). Arrows Localized voids. e Enlarged view<br />

of an area showing the simultaneous degradation (arrow). Arrowheads Autolysing hyphae.<br />

H Sound, turgid hyphae. Bars a 3µm,b 20 µm, c 5µm,d 10 µm, e 2 µm. Reproduced with<br />

permission from Tsuneda et al. 2001


234 A.V. Rice, R.S. Currah<br />

because both consist of cellulose microfibrils embedded in an amorphous<br />

matrix of phenolic polymers and polysaccharides (Tsuneda et al. 2001).<br />

Both species degraded Sphagnum leaves but decay patterns differed, with<br />

O. maius (UAMH 9749) eroding all cell wall components simultaneously<br />

(Fig. 13.2) and P. bulbillosa degrading preferentially the amorphous matrix<br />

material (Tsuneda et al. 2001). This pattern, analogous to the simultaneous<br />

white rot of wood, was confirmed in O. maius and other members of the<br />

Myxotrichaceae isolated from peat (Rice et al. 2006).<br />

These in vitro enzymatic studies, mass loss experiments, and scanning<br />

electron microscopic examinations indicate that O. maius has the potential<br />

to degrade Sphagnum peat in nature. The abundance of O. maius conidia<br />

and conidiophores on peat (Rice and Currah 2002), and the relatively frequent<br />

isolation of O. maius from this material (Barron 1962; Nilsson et al.<br />

1992; Thormann et al. 2001, 2004; Rice and Currah 2002; Rice et al. 2006),<br />

support the hypothesis that O. maius is an active component of the saprobic<br />

microfungal community in peatlands.<br />

13.3<br />

Ericoid Mycorrhizas<br />

Cronquist (1988) recognised eight families within the globally distributed<br />

order Ericales that are integral components of many acidic, nutrient-poor<br />

ecosystems with organic soils. Four families, the Ericaceae, Empetraceae,<br />

Monotropaceae, and Pyrolaceae, are found in the northern hemisphere<br />

(Cronquist 1988). Molecular evidence suggests that these families, along<br />

with the Epacridaceae in the southern hemisphere, should be included<br />

together in the Ericaceae (Kron 1996; Kron et al. 2002). Ericoid and ectendomycorrhizas<br />

are common within the Ericaceae but there are also reports<br />

of ectomycorrhizas (Largent et al. 1980; Smith et al. 1995; Horton et al.<br />

1999) and arbuscular mycorrhizas (Koske et al. 1990).<br />

Most Ericaceae are dwarf shrubs adapted to harsh ecosystems including<br />

bogs, heaths, alpine and arctic regions, and boreal forests (Hambleton<br />

1998). These woody plants have leathery, perennial leaves that minimise<br />

nutrient loss. Ericoid mycorrhizal associations may enhance the success<br />

of host plants in nutrient-poor, acidic, phenol-rich, and heavy metalcontaminated<br />

soils (Perotto et al. 1995; Hambleton 1998; Hambleton and<br />

Currah 2000). The below-ground network consists of well-developed mats<br />

of rhizomes and “hair roots” that form in the surface layers of organic soil<br />

(Read 1991). “Hair roots” have a narrow stele surrounded by an endodermis<br />

and one to two layers of cells, representing the cortex and/or epidermis<br />

(Read 1991; Smith and Read 1997). Hyphae penetrate the cell walls of the<br />

outer cell layers, form an interface with cell membranes (Read 1991; Smith


13 Saprobe and mutualist 235<br />

Fig.13.3. a,b Oidiodendron maius (S. Hambleton personal collection, S-272a) colonising<br />

the roots of Vaccinium vitis-idaea in resynthesis studies (S. Hambleton, unpublished).<br />

a Longitudinal section of mycorrhizal root showing hyphal complexes formed in the outer<br />

layerofrootcells(arrow). b Close up of hyphal complex (arrow) formed in the root cell.<br />

Bars 10 µm. Images provided by Sarah Hambleton<br />

and Read 1997), and develop into complexes made up of densely intertwined,<br />

thin, lightly pigmented hyphae (Fig. 13.3). Hyphae also extend out<br />

of the root and absorb nutrients by decomposing organic matter; some<br />

of these nutrients, at least, are then supplied to the plant (Northup et al.<br />

1995; Smith and Read 1997). Nutrient and carbon exchange is believed<br />

to occur across the interfaces between plant cell membranes and hyphal<br />

complexes for about 5 weeks until both the plant cell cytoplasm and the<br />

fungal hyphae within the cell degenerate (Read 1991; Smith and Read 1997).<br />

Ericoid mycorrhizal fungi have also been shown to break down phenolic<br />

compounds and sequester heavy metal ions, detoxifying the soil for their<br />

plant partner (Read 1991; Perotto et al. 1995; Smith and Read 1997; Yang and<br />

Goulart 2000). While the benefits to the host plant are readily demonstrated<br />

in resynthesis studies, the benefits to the mycobiont (fungal partner) are<br />

more difficult to measure; but it is assumed that the mycobiont receives<br />

photosynthates from the host plant (Smith and Read 1997).<br />

Identification of mycorrhizal fungal symbionts requires isolation and<br />

identification of the fungi in culture, followed by resynthesis of the association<br />

(Smith and Read 1997; Hambleton 1998). Symbiont identification has<br />

relied upon morphological and cultural characteristics of the fungi; these<br />

sources of characters work well in conjunction with DNA fingerprinting<br />

and sequence techniques (e.g. Gardes et al. 1991; Simon et al. 1992; Egger<br />

1995; Clapp et al. 2002; Erland and Taylor 2002; Allen et al. 2003). The<br />

first fungi confirmed, through resynthesis experiments, to form ericoid<br />

mycorrhizas did not sporulate and hence were unidentifiable (Doak 1928;<br />

Bain 1937; Gordon 1937; McNabb 1961). In 1973, Pearson and Read reported<br />

that some of their ericoid mycorrhizal isolates produced zigzag chains of


236 A.V. Rice, R.S. Currah<br />

arthroconidia in culture and one produced small apothecia in pure culture<br />

and in pots containing Calluna vulgaris. The conidial fungus was later<br />

named Scytalidium vaccinii (Dalpé et al. 1989) and the apothecial fungus<br />

was named Pezizella ericae (Read 1974). This species was transferred to<br />

Hymnenoscyphus (Kernan and Finocchio 1983) and recently to Rhizoscyphus,asR.<br />

ericae (Zhang and Zhuang 2004). Scytalidium vaccinii has been<br />

confirmed as the anamorph of R. ericae (Egger and Sigler 1993; Hambleton<br />

1998; Hambleton et al. 1999). Hambleton and Currah (1997) described<br />

a series of isolates under “variable white taxon” (VWT) that were common<br />

endophytes in ericaceous roots. This taxon was shown to have marked<br />

affinities to R. ericae but produced neither a teleomorph nor conidia in culture.<br />

Recently, three phylogenetically distinct species in the VWT complex<br />

have been recognised in a new anamorphic genus (Hambleton and Sigler<br />

2005). Comparison of fungi detected in the roots of Gaultheria shallon using<br />

culturing and molecular (DNA) methods revealed that an abundance of<br />

hyphae within hair roots belonged to an unculturable species of Sebacina<br />

(Allen et al. 2003). Fungi that were culturable included R. ericae and an<br />

isolate tentatively identified as a species of Capronia (Allen et al. 2003). As<br />

detection techniques and identification protocols improve, it is expected<br />

that additional fungal taxa will be described as ericoid mycorrhizal endophytes.<br />

Species of Oidiodendron other than O. maius have also been reported<br />

from ericoid mycorrhizas (Pearson and Read 1973; Couture et al. 1983;<br />

Dalpé 1986, 1989, 1991, Douglas et al. 1989; Xiao and Berch 1992; Currah et<br />

al. 1993, 1999; Johansson 1994, 2001; Perotto et al. 1995, 1996; Hambleton<br />

and Currah 1997; Monreal et al. 1999; Chambers et al. 2000; Usuki et<br />

al. 2003). While many of the early reports implicated O. griseum in the<br />

associations, DNA analyses indicate that most, if not all, of these reports<br />

werebasedonmisidentifiedstrainsofO. maius (Hambleton and Currah<br />

1997; Hambleton et al. 1998).<br />

Mycorrhizal resyntheses between host plants and fungi isolated from<br />

their ericoid mycorrhizas are generally assumed to be the definitive indicator<br />

that a fungus is mycorrhizal, but assumptions based on these<br />

data are tenuous at best. The Ericaceae is particularly problematic in<br />

this regard because, in axenic culture situations at least, the family appears<br />

to permit a wide range of fungi into the peripheral cells of hair<br />

roots where they form the typical coiled “infection units”. For example,<br />

Dalpé (1986, 1989, 1991) found that blueberries (Vaccinium angustifolium)<br />

would form ericoid mycorrhizas with Myxotrichum setosum, O. cerealis,<br />

O. chlamydosporicum, O. citrinum, O. flavum, O. griseum, O. periconioides,<br />

O. rhodogenum, O. scytaloides, Pseudogymnoascus roseus (all members of<br />

the Myxotrichaceae, Leotiomycetes) as well as the unrelated species Gymnascella<br />

dankalienses (a member of the Gymnoascaceae, Eurotiomycetes).


13 Saprobe and mutualist 237<br />

Salal (Gaultheria shallon) has formed in vitro mycorrhizal associations with<br />

Acremonium strictum, a fungus that was able to supply organic nitrogen<br />

and enhance host plant growth (Xiao and Berch 1999). Ericaceous plants<br />

may “prefer” some fungi over others in the field but when constrained, as<br />

in a culture situation, may form mycorrhizal relationships with, or exploit,<br />

a range of different fungal species.<br />

13.4<br />

Oidiodendron maius as an Ericoid Mycorrhizal Fungus<br />

While the widespread isolation of O. maius from the roots of ericaceous<br />

roots supports the hypothesis that it may be mycorrhizal and as such,<br />

either a mutualist or a commensalist, it does not confirm the nature of<br />

the association. Many resynthesis studies have attempted to assess the<br />

morphological and functional aspects of the relationship; these studies have<br />

used a range of ericaceous shrubs and have reported positive, neutral, and<br />

negative effects on host plant growth (Douglas et al. 1989; Xiao and Berch<br />

1995, 1999; Yang et al. 1998; Monreal et al. 1999; Bergero et al. 2000; Yang<br />

and Goulart 2000; Johansson 2001; Starrett et al. 2001; Piercey et al. 2002;<br />

Yang et al. 2002) using a range of Oidiodendron species (Couture et al. 1983;<br />

Dalpé 1986, 1989, 1991; Currah et al. 1993; Xiao and Berch 1995; Monreal<br />

et al. 1999). Characteristic hyphal complexes have been observed in roots<br />

of various ericaceous shrubs, including Vaccinium vitis-idaea, colonized<br />

by Oidiodendron species (Fig. 13.3) (Dalpé 1986, 1989, 1991; Douglas et<br />

al. 1989; Xiao and Berch 1995; Johansson 2001; S. Hambleton, personal<br />

communication). While the plants in these studies appeared healthy, the<br />

functional nature of the relationship between the fungi and the hosts was<br />

not determined.<br />

Physiological evidence to support the mycorrhizal nature of the association<br />

between O. maius and ericaceous plants has been obtained from<br />

resynthesis studies. Yang et al. (1998) found that O. maius (UAMH 9263)<br />

did not affect the growth of blueberries (Vaccinium corymbosum), but later<br />

studies (Yang and Goulart 2000; Yang et al. 2002) on the same isolate of<br />

O. maius and the same plant species found positive effects of the fungus<br />

on plant growth, indicating that the nature of the relationship between<br />

O. maius and ericaceous shrubs may vary within fungal strains. Inoculation<br />

of salal (Gaultheria shallon) withfourisolatesofO. maius increased<br />

plant biomass regardless of the nitrogen source supplied (Xiao and Berch<br />

1999). Inoculation with O. maius increased blueberry root and shoot dry<br />

mass as well as plant access to organic nitrogen (Yang et al. 2000). O. maius<br />

has also been shown to reduce aluminum uptake and increase the cation<br />

exchange capacity of blueberry (Yang and Goulart 2000).


238 A.V. Rice, R.S. Currah<br />

In contrast, several resynthesis studies using O. maius (Dalpé 1991; Bergero<br />

et al. 2000; Piercey et al. 2002) have not resulted in the formation of<br />

distinctive infection units. These results may be due to the strong saprobic<br />

abilities of O. maius, strain specific effects, or to the sources of nutrients<br />

available to the plants. In these axenic culture systems, O. maius might have<br />

acquired sufficient carbon from the substrate (Sphagnum peat in the study<br />

by Piercey et al. 2002) and did not need the host plant for nutrition (Piercey<br />

et al. 2002). Other resynthesis studies have shown either neutral (Yang et<br />

al. 1998) or negative effects on the plants (Starrett et al. 2001). Starrett et al.<br />

(2001) inoculated microshoots of mountain andromeda (Pieris floribunda)<br />

with the “ericoid mycorrhizal fungi” R. ericae, O. maius (ATCC 66504),<br />

O. griseum and an unidentified species of Oidiodendron, and found that inoculation<br />

with all of these species caused shoot necrosis, but that this effect<br />

could be reduced by providing an alternative carbon source. Mitigation of<br />

shoot necrosis varied with carbon source and fungal isolate. Adding sucrose<br />

to the medium prevented R. ericae,butnottheOidiodendron species, from<br />

causing shoot necrosis while adding a peat-vermiculite mixture reduced<br />

the shoot necrosis caused by the Oidiodendron species. Additionally, the<br />

Oidiodendron species did not induce root formation by the microshoots to<br />

the same extent as R. ericae, leading Starrett et al. (2001) to conclude that<br />

the Oidiodendron species were not mycorrhizal symbionts.<br />

None of the preceding studies investigated possible benefits to the<br />

mycobiont, so the mutualistic nature of the association has never been<br />

demonstrated. It is possible that the relationship is physiologically similar<br />

to the dynamics in orchid mycorrhizas in which the orchid “exploits”<br />

its saprobic or ectomycorrhizal mycobiont (e.g. Rasmussen 1995;<br />

McKendrick et al. 2000) or to the epiparasitic relationship of monotropes<br />

(non-photosynthetic Ericaceae) to ectomycorrhizal trees via their shared<br />

ectomycorrhizal fungi (Bidartondo et al. 2000). Many Ericaceae, similar to<br />

the Orchidaceae and the monotropes, are microspermous (e.g. Rhododendron,<br />

Menziesia). Perhaps the adoption of microspermy is a consequence of<br />

the host plants’ ability to exploit fungi as a source of nutrients, and ericoid<br />

mycorrhizal associations may be a part of a mycoheterotrophic evolutionary<br />

trajectory. The discovery of Sebacina, a basidiomycete genus known to<br />

form orchid mycorrhizas (Currah et al. 1990; McKendrick et al. 2002), in<br />

ericoid mycorrhizal roots of Salal (Allen et al. 2003) is another similarity<br />

between ericoid and orchid mycorrhizal systems.<br />

The nature of the relationship between O. maius and members of the<br />

Ericaceae is clearly complex and varies from one report to the next. Future<br />

research involving physiological assessment in laboratory, greenhouse, and<br />

field conditions coupled with morphological and molecular assessment in<br />

roots (Hambleton and Currah 2000) is required to identify and quantify<br />

possible benefits to both partners and to elucidate the factors determining


13 Saprobe and mutualist 239<br />

the functional aspects of the relationship. For example, tracing the movement<br />

of radiolabeled carbon and nutrients between the partners could help<br />

determine whether O. maius obtains host photosynthate and if the plant<br />

receives any fungal-derived carbon or other nutrients. Additionally, in vitro<br />

studies could determine other potential benefits to either partner, such as<br />

competitive and growth advantages for O. maius and pathogen resistance<br />

for the plant. The potential role of ericoid mycorrhizal fungi in symbiotic<br />

seed germination should also be examined.<br />

13.5<br />

Significance and Relevance<br />

Theoccupationofmultiplenicheswithinagivenenvironmentcouldconfer<br />

survival and competitive advantages on O. maius by providing different<br />

refuges to the fungus. During periods of host plant dormancy, O. maius<br />

could thrive as a saprobe in the peat matrix, while host plant roots may<br />

serve as a refuge from competition with other saprobes and as a source of<br />

inoculum for colonisation of senescing surface peat. On the other hand,<br />

the prevalence of O. maius as a saprobe within the peat could ensure<br />

rapid colonisation of new ericaceous roots, perhaps at the expense of other<br />

species that are less able to degrade the surrounding substrate. While there<br />

is currently no experimental data to support either hypothesis (i.e. whether<br />

roots or peat serve as primary refugia for O. maius), testing both could<br />

provide information key to understanding of the ecological role of this<br />

species.<br />

Ericoid and ectomycorrhizal fungi can degrade a variety of complex<br />

organic substrates (Bajwa and Read 1985; Read 1991; Northup et al. 1995;<br />

Bending and Read 1996, 1997; Smith and Read 1997; Aerts 2002; Leake et al.<br />

2002; Olsson et al. 2002; Simard et al. 2002), with the abilities of ericoid mycorrhizal<br />

fungi possibly exceeding those of ectomycorrhizal fungi (Bajwa<br />

and Read 1985; Read 1991; Bending and Read 1996, 1997; Smith and Read<br />

1997). These abilities are thought to aid in host plant nutrition by allowing<br />

the plant direct access to organic nutrient sources (Bajwa and Read 1985;<br />

Xiao and Berch 1999; Yang et al. 2002). Inorganic sources of nitrogen are<br />

scarce and organic sources relatively abundant when decomposition is slow,<br />

as in peatlands and heathlands, and when leaching is common (Perotto et<br />

al. 1995). Ericoid mycorrhizal fungi often produce phosphatase, enabling<br />

them to access and transfer organic phosphorus to their hosts (Aerts 2002).<br />

Ericaceous shrubs in these environments may rely on their mycorrhizal<br />

partners to supply them with sufficient carbon (Yang et al. 2002) and nutrients<br />

obtained from the organic sources (Northup et al. 1995; Xiao and Berch<br />

1999; Yang et al. 2002). The abilities of ericoid mycorrhizal fungi, including


240 A.V. Rice, R.S. Currah<br />

O. maius, to access carbon and nutrients from organic debris could reduce<br />

their reliance on host plant photosynthates for carbon (Piercey et al. 2002),<br />

while still supplying the host plant with nutrients.<br />

Transfer of nutrients from organic matter to ericaceous shrubs via ericoid<br />

mycorrhizal fungi, such as O. maius, has important implications for<br />

nutrient cycling in ecosystems such as peatlands, where decomposition is<br />

slow and carbon and nutrients are sequestered in organic debris (Northup<br />

et al. 1995). Sphagnum decomposes slowly with relatively few fungi having<br />

the ability to cause significant mass losses (Thormann 2001). Decomposition<br />

of Sphagnum by O. maius may release a significant amount of carbon<br />

and nutrients from the peat and, instead of releasing carbon into the atmosphere<br />

and nitrogen and phosphorus into the pool of plant-available<br />

nutrients, these organic forms of nitrogen and phosphorus can be supplied<br />

directly to the ericaceous shrubs, giving them a competitive advantage over<br />

their neighboring plants (Aerts 2002; Leake et al. 2002). Short-circuiting<br />

nutrient cycles, by reducing the amount of decomposition required before<br />

nutrient absorption, results in increased supplies of nitrogen and phosphorus<br />

to mycorrhizal host plants and a resultant decrease in nitrogen and<br />

phosphorus available to saprobes and non-mycorrhizal plants (Leake et al.<br />

2002).<br />

Other Oidiodendron species are able to form ericoid mycorrhizal associations<br />

in vitro (Dalpé 1986, 1989, 1991; Currah et al. 1993). Species<br />

of Oidiodendron are the asexual states of myxotrichoid ascomycetes, and<br />

some of these, e.g. Pseudogymnoascus roseus, and other anamorphs, including<br />

species of Geomyces, also produce ericoid mycorrhizal associations in<br />

vitro (Dalpé 1989). However, only two species of Oidiodendron have been<br />

reported from ericaceous roots in situ. Currah et al. (1993) isolated O. periconioides<br />

from Rhododendron brachycarpum growninpotculturescontaining<br />

peat. The remaining species are known only as saprobes but since<br />

they share morphological characters, including dendritic arthroconidia<br />

and cage-like cleistothecial ascomata, and ecological characters, including<br />

the ability to degrade a variety of plant-based polymers and a predilection<br />

for cool, acidic conditions (Rice and Currah 2005), it is possible that these<br />

taxa could play biologically similar and significant roles to O. maius in cool,<br />

acidic soils. Additional surveys of ericoid mycorrhizal endophytes should<br />

employ a broad range of isolation and detection protocols to maximise the<br />

recovery of as wide a variety of fungi as possible.<br />

The occupation of multiple niches by O. maius may parallel the situation<br />

observed for Phialocephala fortinii. Usuallyconsideredarootendophyte,<br />

P. fortinii is isolated most frequently from healthy roots of woody plants [e.g.<br />

see Chaps. 7 (Sieber and Grünig) and 15 (Schulz); Stoyke and Currah 1991;<br />

Menkis et al. 2004; Piercey et al. 2004] and, until recently, was unknown as<br />

a saprobe. However, Menkis et al. (2004) isolated P. fortinii from healthy


13 Saprobe and mutualist 241<br />

wood in pine stems suggesting a possible role as a systemic endophyte, and<br />

inbirchsnagsandbirchandpinestumps,whereitispresumablyoccupying<br />

a saprobic niche (Menkis et al. 2004), perhaps as an agent of soft rot (Sieber<br />

2002; Menkis et al. 2004). Given the scattered and somewhat incidental<br />

reports of O. maius from wood and soil, a concerted and wider search for<br />

O. maius may reveal habitats in addition to ericaceous roots, where this<br />

speciesisabundant.<br />

13.6<br />

Conclusions<br />

Oidiodendron maius formsassociationswiththerootsofericaceousshrubs,<br />

though the nature of the relationship remains uncertain. Is it a mutualistic<br />

mycorrhizal association, a preemptively colonised refugium for the fungus,<br />

a case of parasitism of the fungus by the plant, or some combination of<br />

the three? In vitro studies indicate that O. maius can improve host plant<br />

growth both by aiding plant nutrition and detoxifying the soil environment,<br />

although the benefits to O. maius are unclear. It remains necessary<br />

to investigate the benefits to both partners and demonstrate what environmental<br />

conditions determine the functional nature of the relationship.<br />

O. maius has the potential to degrade complex organic polymers within<br />

the soil, thus it is unlikely that it would rely on host photosynthate for<br />

survival. However, it is possible that O. maius receives some photosynthate,<br />

which could supplement saprobically derived carbon, potentially giving<br />

O. maius a competitive advantage over other soil fungi. The tendency towards<br />

microspermy in the Ericaceae and the saprobic abilities of ericoid<br />

endophytes suggests that ericoid mycorrhizal associations may represent<br />

another example of controlled parasitism of a fungal partner by the host<br />

plant, similar to the type that occurs with orchids. Entrapment of O. maius<br />

could confer a competitive advantage on the host plants by increasing the<br />

supply of organically bound nutrients-unavailable to plants that lack ericoid<br />

mycorrhizas, and by supplementing host plant photosynthesis with<br />

fungal-derived carbon. Given the wide taxonomic tolerance that ericaceous<br />

plants have for root endophytic fungi in vitro, the obvious need for more<br />

detailed studies of endophytic diversity of fungi growing in plants in situ,<br />

the enigmatic ecological roles of related Helotialean fungi (e.g. P. fortinii,<br />

Geomyces, etc.), much more exploratory and empirical research is needed<br />

before we will be able to answer the question posed at the outset of this<br />

chapter.


242 A.V. Rice, R.S. Currah<br />

Acknowledgements. The authors thank S. Hambleton and A. Tsuneda for<br />

providing images. Comments on previous versions of this manuscript by<br />

H.D. Addy and the continuing support of the Natural Sciences and Engineering<br />

Research Council of Canada are gratefully acknowledged.<br />

<strong>References</strong><br />

Addy HD, Piercey MM, Currah RS (2005) Microfungal endophytes in roots. Can J Bot 83:1–13<br />

Aerts R (2002) The role of various types of mycorrhizal fungi in nutrient cycling and plant<br />

competition. In: van der Heijden MGA, Sanders IR (eds) Ecological studies, vol 157<br />

Mycorrhizal ecology Springer, Berlin Heidelberg New York, pp 117–133<br />

Allen TR, Millar T, Berch SM, Berbee ML (2003) Culturing and direct DNA extraction find<br />

different fungi from the same ericoid mycorrhizal roots. New Phytol 160:255–272<br />

Bain HF (1937) Production of synthetic mycorrhiza in the cultivated cranberry. J Agric Res<br />

55:811–835<br />

Bajwa R, Read DJ (1985) The biology of mycorrhiza in the Ericaceae. IX. Peptides as nitrogen<br />

sources for the ericoid endophyte and for mycorrhizal and non-mycorrhizal plants. New<br />

Phytol 101:459–467<br />

Barron GL (1962) New species and new records of Oidiodendron. Can J Bot 40:589–607<br />

Bending GD, Read DJ (1996) Nitrogen mobilization from tannin-protein complexes by<br />

ericoid and ectomycorrhizal fungi. Soil Biol Biochem 28:1603–1612<br />

Bending GD, Read DJ (1997) Lignin and soluble phenolic degradation by ectomycorrhizal<br />

and ericoid mycorrhizal fungi. Mycol Res 101:1348–1354<br />

Bergero R, Perotto S, Girlanda MM, Vidano G, Luppi AM (2000) Ericoid mycorrhizal fungi<br />

are common root associates of a Mediterranean ectomycorrhizal plant (Quercus ilex).<br />

Mol Ecol 9:1639–1649<br />

Bidartondo MI, Kretzer AM, Pine EM, Bruns TD (2000) High root concentrations and<br />

uneven ectomycorrhizal diversity near Sarcodes sanguinea (Ericaceae): A cheater that<br />

stimulates its victims? Am J Bot 87:1783–1788<br />

Chambers SM, Liu G, Cairney WG (2000) ITS rDNA sequence comparison of ericoid mycorrhizal<br />

endophytes from Woollsia pungens. Mycol Res 104:168–174<br />

Clapp JP, Helgason T, Daniell TJ, Young JPW (2002) Genetic studies of the structure and<br />

diversity of arbuscular mycorrhizal fungal communities. In: van der Heijden MGA,<br />

Sanders IR (eds) Ecological studies, vol 157. Mycorrhizal ecology Springer, Berlin Heidelberg<br />

New York, pp 201–224<br />

Couture M, Fortin JA, Dalpé Y (1983) Oidiodendron griseum Robak: an endophyte of ericoid<br />

mycorrhiza in Vaccinium spp. New Phytol 95:375–380<br />

Cronquist A (1988) The evolution and classification of flowering plants, 2nd edn. New York<br />

Botanic Garden, New York<br />

Currah RS, Smreciu ES, Hambleton S (1990) Mycorrhizae and mycorrhizal fungi of boreal<br />

species of Platanthera and Coeloglossum (Orchidaceae). Can J Bot 68:1171–1181<br />

Currah RS, Tsuneda A, Murakami S (1993) Conidiogenesis in Oidiodendron periconioides<br />

and ultrastructure of ericoid mycorrhizas formed with Rhododendron brachycarpum.<br />

Can J Bot 71:1481–1485<br />

Currah RS, Niemi M, Huhtinen S (1999) Oidiodendron maius and Scytalidium vaccinii from<br />

the mycorrhizas of Ericaceae in northern Finland. Karstenia 39:65–68<br />

Dalpé Y (1986) Axenic synthesis of ericoid mycorrhiza in Vaccinium angustifolium Ait. by<br />

Oidiodendron species. New Phytol 103:391–396


13 Saprobe and mutualist 243<br />

Dalpé Y (1989) Ericoid mycorrhizal fungi in the Myxotrichaceae and Gymnoascaceae. New<br />

Phytol 113:523–527<br />

Dalpé Y (1991) Statut endomycorhizien du genre Oidiodendron. Can J Bot 69:1712–1714<br />

Dalpé Y, Litten W, Sigler L (1989) Scytalidium vaccinii sp. nov., an ericoid endophyte of<br />

Vaccinium angustifolium roots. Mycotaxon 35:371–377<br />

Doak KD (1928) The mycorrhizal fungus of Vaccinium. Phytopathology 18:148<br />

Douglas AE, Smith DC (1989) Are endosymbioses mutualistic. Tree 4:350–352<br />

Douglas GC, Heslin MC, Reid C (1989) Isolation of Oidiodendron maius from Rhododendron<br />

and ultrastructural characterization of synthesized mycorrhizas. Can J Bot 67:2206–2212<br />

Egger KN (1995) Molecular analysis of ecto-mycorrhizal fungal communities. Can J Bot<br />

73:S1415–S1422<br />

Egger KN, Sigler L (1993) Relatedness of the ericoid endophytes Scytalidium vaccinii and<br />

Hymenoscyphus ericae inferred from analysis of ribosomal DNA. Mycologia 85:219–230<br />

Erland S, Taylor AFS (2002) Diversity of ecto-mycorrhizal fungal communities in relation to<br />

the abiotic environment. In: van der Heijden MGA, Sanders IR (eds) Ecological studies,<br />

vol 157. Mycorrhizal ecology Springer, Berlin Heidelberg New York, pp 163–200<br />

Gardes M, White TJ, Fortin JA, Bruns TD, Taylor JW (1991) Identification of indigenous<br />

and introduced symbiotic fungi in ecto-mycorrhizas by amplification of nuclear and<br />

mitochondrial ribosomal DNA. Can J Bot 69:180–190<br />

Gordon HD (1937) Mycorrhiza in Rhododendron. Ann Bot 1:593–613<br />

Hambleton S (1998) Mycorrhizas of the Ericaceae: Diversity and systematics of the mycobionts.<br />

PhD Dissertation, University of Alberta, Edmonton<br />

Hambleton S, Currah RS (1997) Fungal endophytes from the roots of alpine and boreal<br />

Ericaceae. Can J Bot 75:1570–1581<br />

Hambleton S, Currah RS (2000) Molecular characterization of the mycorrhizas of woody<br />

plants. In: Jain SM, Minocha SC (eds) Molecular biology of woody plants, vol 2. Kluwer,<br />

Dordrecht, pp 351–373<br />

Hambleton S, Sigler L (2005) Melinia, a new anamorph genus for root-associated fungi with<br />

phylogenic affinities to Rhizoscyphus ericae (≡ Hymenoscyphus ericae), Leotiomycetes.<br />

Stud Mycol 53:1–27<br />

Hambleton S, Egger KN, Currah RS (1998) The genus Oidiodendron: species delimitation<br />

and phylogenetic relationships based on nuclear ribosomal DNA analyses. Mycologia<br />

90:854–869<br />

Hambleton S, Huhtinen S, Currah RS (1999) Hymenoscyphus ericae: a new record from<br />

western Canada. Mycol Res 103:1391–1397<br />

Hoosbeek M, van Breemen N, Berendse F, Grosvernier P, Vasander H, Wallén B (2001) Limited<br />

effect of increased atmospheric CO2 concentration on ombrotrophic bog vegetation.<br />

New Phytol 150:459–463<br />

Horton TR, Bruns TD, Parker VT (1999) Ectomycorrhizal fungi associated with Arctostaphylos<br />

contribute to Pseudotsuga menziesii establishment. Can J Bot 77:93–102<br />

Hutchison LJ (1990) Studies on the systematics of ectomycorrhizal fungi in axenic culture.<br />

II. The enzymatic degradation of selected carbon and nitrogen compounds. Can J Bot<br />

68:1522–1530<br />

Hutchison LJ (1991) Description and identification of cultures of ectomycorrhizal fungi<br />

found in North America. Mycotaxon 42:387–504<br />

Johansson M(1994) Quantification of mycorrhizal infection units in roots ofCalluna vulgaris<br />

(L.) Hull from Danish heathland. Soil Biol Biochem 26:557–566<br />

Johansson M (2001) Fungal associations of Danish Calluna vulgaris roots with special<br />

reference to ericoid mycorrhiza. Plant Soil 231:225–232<br />

Kernan MJ, Finocchio AF (1983) A new discomycete associated with the roots of Monotropa<br />

uniflora (Ericaceae). Mycologia 75:916–920


244 A.V. Rice, R.S. Currah<br />

Koske RE, Gemma JN, Englander L (1990) Vesicular-arbuscular mycorrhizae in Hawaiian<br />

Ericales. Am J Bot 77:64–68<br />

Kron KA (1996) Phylogenetic relationships of Empetraceae, Epacridaceae, Ericaceae,<br />

Monotropaceae, and Pyrolaceae: evidence from nuclear ribosomal 18S sequence data.<br />

Ann Bot 77:293–303<br />

Kron KA, Judd WS, Stevens PF, Crayne DM, Anderberg AA, Gadek PA, Quinn CJ, Luteyn JL<br />

(2002) Phylogenetic classification of Ericaceae: molecular and morphological evidence.<br />

Bot Rev 68:335–423<br />

Largent DL, Sugihara N, Wishner C (1980) Occurrence of mycorrhizae on ericaceous and<br />

pyrolaceous plants in northern California. Can J Bot 58:2274–2279<br />

Leake JR, Donnelly DP, Boddy L (2002) Interactions between ecto-mycorrhizal and saprotrophic<br />

fungi. In: van der Heijden MGA, Sanders IR (eds) Ecological studies, vol 157.<br />

Mycorrhizal Ecology. Springer, Berlin Heidelberg New York, pp 345–372<br />

Lumley TC, Gignac LD, Currah RS (2001) Microfungus communities of white spruce and<br />

trembling aspen logs at different stages of decay in disturbed and undisturbed sites in<br />

the boreal mixedwood region of Alberta. Can J Bot 79:76–92<br />

McKendrick SL, Leake JR, Read DJ (2000) Symbiotic germination and development of mycoheterotrophic<br />

plants in nature: transfer of carbon from ectomycorrhizal Salix repens and<br />

Betula pendula to the orchid Corallorhiza trifida through shared hyphal connections.<br />

New Phytol 145:539–548<br />

McKendrick SL, Leake JR, Taylor DL, Read DJ (2002) Symbiotic germination and development<br />

of the myco-heterotrophic orchid Neottia nidus-avis in nature and its requirement<br />

for locally distributed Sebacina. New Phytol 154:233–247<br />

McNabb RFR (1961) Mycorrhiza in the New Zealand Ericales. Aust J Bot 9:57–61<br />

Menkis A, Allmer J, Vasiliauskas R, Lygis V, Stenlid J, Finlay R (2004) Ecology and molecular<br />

characterization of dark septate fungi in roots, living stems, coarse and fine woody<br />

debris. Mycol Res 108:965–973<br />

Monreal M, Berch SM, Berbee M (1999) Molecular diversity of ericoid mycorrhizal fungi.<br />

Can J Bot 77:1580–1594<br />

Nilsson M, Bååth E, Söderström B (1992) The microfungal communities of a mixed mire in<br />

Northern Sweden. Can J Bot 70:272–276<br />

Nordgren A, Bååth E, Söderström B (1985) Soil microfungi in an area polluted by heavy<br />

metals. Can J Bot 63:448–455<br />

Northup RR, Zengshou Y, Dahlgren RA, Vogt KA (1995) Polyphenol control of nitrogen<br />

release from pine litter. Nature 377:227–229<br />

Olsson PA, Jakobsen I, Wallander H (2002) Foraging and resource allocation strategies of<br />

mycorrhizal fungi in a patchy environment. In: van der Heijden MGA, Sanders IR (eds)<br />

Ecological studies, vol 157. Mycorrhizal ecology Springer, Berlin Heidelberg New York,<br />

pp 93–115<br />

Pearson V, Read DJ (1973) The biology of mycorrhiza in Ericaceae. I. The isolation of the<br />

endophyte and synthesis of mycorrhizas in aseptic culture. New Phytol 72:371–379<br />

Perotto S, Perotto R, Faccio A, Schubert A, Varma A, Bonfante P (1995) Ericoid mycorrhizal<br />

fungi: cellular and molecular bases of their interactions with the host plant. Can J Bot<br />

73:S557–S568<br />

Perotto S, Actis-Perino E, Perugini J, Bonfante P (1996) Molecular diversity of fungi from<br />

ericoid mycorrhizal roots. Mol Ecol 5:123–131<br />

Peteet D, Andreev A, Bardeen W, Mistretta F (1998) Long-term arctic peatland dynamics,<br />

vegetation and climate history of the Pur-Taz region, western Siberia. Boreas 27:115–126<br />

Piercey MM, Thormann MN, Currah RS (2002) Saprobic characteristics of three fungal<br />

taxa from ericalean roots and their association with the roots of Rhododendron groenlandicum<br />

and Picea mariana in culture. Mycorrhiza 12:175–180


13 Saprobe and mutualist 245<br />

Piercey MM, Graham SW, Currah RS (2004) Patterns of genetic variation in Phialocephala<br />

fortinii across a broad latitudinal transect in Canada. Mycol Res 108:955–964<br />

Qian XM, El-Ashker A, Kottke I, Oberwinkler F (1998) Studies of pathogenic and antagonistic<br />

microfungal populations and their potential interactions in the mycorrhizoplane<br />

of Norway spruce (Picea abies (L.) Karst.) and beech (Fagus sylvatica L.) on acidified<br />

and limed plots. Plant Soil 199:111–116<br />

Rasmussen H (1995) Terrestrial orchids: From seed to mycotrophic plant. Cambridge University<br />

Press, Cambridge<br />

Read DJ (1974) Pezizella ericae sp. nov., the perfect state of a typical mycorrhizal endophyte<br />

of Ericaceae. Trans Br Mycol Soc 63:381–383<br />

Read DJ (1991) Mycorrhizas in ecosystems. Experientia 47:376–391<br />

Read DJ (2002) Towards ecological relevance – progress and pitfalls in the path towards an<br />

understanding of mycorrhizal functions in nature. In: van der Heijden MGA, Sanders IR<br />

(eds) Ecological studies, vol 157. Mycorrhizal ecology. Springer, Berlin Heidelberg New<br />

York, pp 1–29<br />

Rice AV, Currah RS (2001) Physiological and morphological variation in Oidiodendron<br />

maius. Mycotaxon 79:383–396<br />

Rice AV, Currah RS (2002) New perspectives on the niche and holomorph of the myxotrichoid<br />

hyphomycete, Oidiodendron maius. Mycol Res 106:1463–1467<br />

Rice AV, Currah RS (2005) Oidiodendron: a survey of the named species and unnamed<br />

anamorphs of Myxotrichum. Stud Mycol 53:83–120<br />

Rice AV, Tsuneda A, Currah RS (2006) In vitro decomposition of Sphagnum by some<br />

microfungi resembles white rot of wood. FEMS Microbiol Ecol 56:372–382<br />

Schild DE, Kennedy A, Stuart MR (1988) Isolation of symbiont and associated fungi from<br />

ectomycorrhizas of Sitka spruce. Eur J Forest Pathol 18:51–61<br />

Seiber TN (2002) Fungal root endophytes. In Wasel Y, Eshel A, Kafkafi U (eds) Plant roots:<br />

the hidden half. Dekker, New York, pp 887–917<br />

Simard SW, Jones MD, Durrall DM (2002) Carbon and nutrient fluxes within and between<br />

mycorrhizal plants. In: van der Heijden MGA, Sanders IR (eds) Ecological studies, vol<br />

157. Mycorrhizal ecology Springer, Berlin Heidleberg New York, pp 33–74<br />

Simon L, Lalonde M, Bruns TD (1992) Specific amplification of the 18S fungal ribosomal<br />

genes from VA endomycorrhizal fungi colonizing roots. Appl Environ Microbiol<br />

58:291–295<br />

Smith JE, Molina R, Perry D (1995) Occurrence of ectomycorrhizas on ericaceous and<br />

coniferous seedlings grown in soils from the Oregon Coast Range. New Phytol 129:73–81<br />

Smith SE, Read DJ (1997) Mycorrhizal symbioses, 2nd edn. Academic Press, San Diego<br />

Starrett MC, Blazich FA, Shafer SR, Grand LF (2001) In vitro colonization of micropropagated<br />

Pieris floribunda by ericoid mycorrhizae. I. Establishment of mycorrhizae on<br />

microshoots. HortScience 36:353–356<br />

Stoyke G, Currah RS (1991) Endophytic fungi from the mycorrhizae of alpine ericoid plants.<br />

Can J Bot 69:347–352<br />

Summerbell RC (1987) Microfungi associated with the mycorrhizal mantle and adjacent<br />

microhabitats within the rhizosphere of black spruce. Can J Bot 65:1085–1095<br />

Svensson BM (1995) Competition between Sphagnum fuscum and Drosera rotundifolia:<br />

a case of ecosystem engineering. Oikos 74:205–212<br />

Thormann MN (2001) The fungal communities of decomposing plants in southern boreal<br />

peatlands of Alberta, Canada. PhD Dissertation, University of Alberta, Edmonton<br />

Thormann MN, Currah RS, Bayley SE (2001) Microfungi isolated from Sphagnum fuscum<br />

from a southern boreal bog in Alberta, Canada. Bryologist 104:548–559<br />

Thormann MN, Currah RS, Bayley SE (2002) The relative ability of fungi from Sphagnum<br />

fuscum to decompose selected carbon substrates. Can J Microbiol 48:204–211


246 A.V. Rice, R.S. Currah<br />

Thormann MN, Currah RS, Bayley SE (2004) Patterns of distribution of microfungi in<br />

decomposing bog and fen plants. Can J Bot 82:710–720<br />

Tsuneda A, Currah RS (2004) Ascomatal morphogenesis in Myxotrichum arcticum supports<br />

the derivation of the Myxotrichaceae from a discomycetous ancestor. Mycologia<br />

96:627–635<br />

Tsuneda A, Thormann MN, Currah RS (2001) Modes of cell wall degradation of Sphagnum<br />

fuscum by Acremonium cf. curvulum and Oidiodendron maius. Can J Bot 79:93–100<br />

Usuki F, Abe JP, Kakishima M (2003) Diversity of ericoid mycorrhizal fungi isolated from<br />

hair roots of Rhododendron obtusum var. kaempferi in a Japanese red pine forest.<br />

Mycoscience 44:97–102<br />

Vitt DH, Halsey LA, Thormann MN, Martin T (1996) Peatland inventory of Alberta Phase 1:<br />

overview of peatland resources in the natural regions and subregions of the province.<br />

Alberta Peatland Resource Centre, Edmonton, Publication 96–1<br />

Vrålstad T, Fossheim T, Schumacher T (2000) Piceirhiza bicolorata – the ectomycorrhizal<br />

expression of the Hymenoscyphus ericae aggregate? New Phytol 145:549–563<br />

Vrålstad T, Schumacher T, Taylor AFS (2002) Mycorrhizal synthesis between fungal strains<br />

of the Hymenoscyphus ericae aggregate and potential ectomycorrhizal and ericoid hosts.<br />

New Phytol 153:143–152<br />

Xiao G, Berch S (1992) Ericoid mycorrhizal fungi of Gaultheria shallon. Mycologia<br />

84:470–471<br />

Xiao G, Berch S (1995) The ability of known ericoid mycorrhizal fungi to form mycorrhizae<br />

with Gaultheria shallon on forest clearcuts. Mycologia 87:467–470<br />

Xiao G, Berch S (1999) Organic nitrogen use by salal ericoid mycorrhizal fungi from northern<br />

VancouverIslandandimpactsongrowthinvitroofGaultheria shallon. Mycorrhiza<br />

9:145–149<br />

Yang WQ, Goulart BL (2000) Mycorrhizal infection reduces short-term aluminum uptake<br />

and increases root cation exchange capacity of highbush blueberry plants. HortScience<br />

35:1083–1086<br />

Yang WQ, Goulart BL, Demchak K (1998) Mycorrhizal infection and plant growth of highbush<br />

blueberry in fumigated soil following soil amendment and inoculation with mycorrhizal<br />

fungi. HortScience 33:1136–1137<br />

Yang WQ, Goulart BL, Demchak K, Li Y (2002) Interactive effects of mycorrhizal inoculation<br />

and organic soil amendments on nitrogen acquisition and growth of highbush blueberry.<br />

J Am Soc Hortic Sci 127:742–748<br />

Zhang Y-H, Zhuang W-Y (2004) Phylogenetic relationships of some members in the genus<br />

Hymenoscyphus (Ascomycetes, Helotiales). Nova Hedwigia 78:475–484


14<br />

14.1<br />

Introduction<br />

Mycorrhizal and Endophytic Fungi<br />

of Epacrids (Ericaceae)<br />

John W.G. Cairney<br />

Epacrids are a group of over 450 species of woody plants that were conventionally<br />

classified as the family Epacridaceae (Powell et al. 1996). Detailed<br />

phylogenetic analyses based on morphological and molecular data indicate<br />

that epacrids represent a lineage within Ericaceae, with a recent classification<br />

regarding epacrids as Styphelioideae, one of eight Ericaceae subfamilies<br />

(Kron et al. 2002). Although epacrids occur in several southern hemisphere<br />

locations, including New Zealand, south east Asia, Pacific Ocean<br />

Islands and Patagonia, they are primarily an Australian group (Copeland<br />

1954; Powell et al. 1996). Species richness is greatest in Western Australia<br />

(WA), with some 181 named species occurring in this state, 98% of which<br />

are endemic to WA (Keighery 1996). Seven epacrid tribes are currently<br />

recognised (Kron et al. 2002), most of which comprise small-to-medium<br />

sized shrubby heath-like plants. Some Richeeae taxa, however, resemble<br />

large arborescent monocots (Allaway 1996). Epacrids occupy a range of<br />

habitats that includes dry sandy heathlands and sclerophyll forests, along<br />

with wet alpine bogs and Magellanic tundra (Read 1996). They generally<br />

occur in acidic or neutral soils, but are occasionally found in more basic<br />

soils (Keighery 1996). Despite being geographically and hydrologically<br />

disparate, these habitats characteristically encompass soils in which availability<br />

of mineral nutrients is relatively poor (Read 1996).<br />

In common with many other Ericaceae taxa, epacrids produce extremely<br />

finelateralrootsknownashairroots.Uptothreeordersofhairrootsmaybe<br />

present, and their structure is broadly similar in all taxa: a stele, surrounded<br />

by two layers of suberised cortical cells and an epidermis. Hair roots lack<br />

root hairs, but the surface is generally covered by a mucilage layer (Cairney<br />

and Ashford 2002). When collected from the field, a proportion of the<br />

epidermal cells of epacrid hair roots invariably contains hyphal structures<br />

that are morphologically similar to those produced by ericoid mycorrhizal<br />

John W.G. Cairney: Centre for Plant and Food Sciences, Parramatta Campus, University<br />

of Western Sydney, Locked Bag 1797, Penrith South DC, NSW 1797, Australia, E-mail:<br />

j.cairney@usw.edu.au<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


248 J.W.G. Cairney<br />

fungi in roots of other Ericaceae subfamilies (Reed 1989, 1996). Up to 90%<br />

of the hair root length can display this type of colonisation (Davies et al.<br />

2003), but in most cases considerably less root length is colonised and there<br />

is evidence that colonisation varies on a seasonal basis (Reed 1996; Hutton<br />

et al. 1994; Davies et al. 2003; Kemp et al. 2003).<br />

Detailed analysis of mycorrhiza-like infections of epacrid hair roots generally<br />

reveals longitudinally oriented surface hyphae from which perpendicular<br />

branches arise. Penetration of individual epidermal cells is generally<br />

by a single hypha, sometimes involving an appressorium, and is followed<br />

by development of a hyphal coil in the periplasmic space (Cairney and Ashford<br />

2002). These features, together with the fact that the fungi are generally<br />

ascomycetes, as evidenced by the presence of simple septa and Woronin<br />

bodies (Allen et al. 1989; Steinke et al. 1996; Briggs and Ashford 2001), are<br />

shared with ericoid mycorrhizal infection of other Ericaeae, and are generally<br />

taken to indicate mycorrhizal infection. Aside from putatively ericoid<br />

mycorrhizal ascomycetes, arbuscular mycorrhiza-like infections, suggesting<br />

the presence of Glomales taxa, have also been observed as apparent<br />

endophytes of field-collected epacrid hair roots (Khan 1978; McGee 1986;<br />

Bellgard 1991; McLean and Lawrie 1996; Reed 1996; Davies et al. 2003),<br />

and there is a single report of basidiomycete hyphae in epidermal cells of<br />

an epacrid (Allen et al. 1989). As emphasised by Reed (1996), however, the<br />

presence of infection does not necessarily indicate a mutualistic association.<br />

In the case of apparent arbuscular mycorrhizal infection at least, it<br />

seems most plausible that this reflects opportunistic infection by arbuscular<br />

mycorrhizal fungal hyphae that grew from neighbouring non-Ericaceae<br />

hosts (Reed 1996).<br />

14.2<br />

Endophytes of Epacrid Roots<br />

To date, only some of the endophytes obtained from Australian epacrid<br />

roots have been tested for their abilities to form ericoid mycorrhizas with<br />

host plants and, where this has been conducted, mycorrhizal status has<br />

been inferred simply from production of mycorrhiza-like coils in epidermal<br />

cells. For this reason these endophytes are referred to as putative ericoid<br />

mycorrhizal fungi throughout this chapter.<br />

Many endophytes have been isolated from surface-sterilised epacrid<br />

roots into axenic agar culture. These are typically isolated as sterile mycelia,<br />

rendering their classification on the basis of morphology difficult. Nonetheless,<br />

studies that grouped isolates on the basis of gross morphological<br />

characteristics of cultured mycelia and/or by pectic zymogram patterns<br />

established that multiple ascomycete taxa are likely to form ericoid mycor-


14 Mycorrhizal and Endophytic Fungi of Epacrids (Ericaceae) 249<br />

rhizal associations with epacrids in their native habitats (Reed 1989; Hutton<br />

et al. 1994, 1996b; Steinke et al. 1996). Subsequent molecular analyses of the<br />

isolated endophytes suggest that they are broadly similar to those obtained<br />

from other Ericaceae taxa in northern hemisphere habitats.<br />

To date, molecular analysis of endophytes from epacrid roots has concentrated<br />

largely on the rDNA internal transcribed spacer (ITS) region.<br />

Assemblages of isolated endophytes have thus been grouped on the basis<br />

of gross morphological characteristics of cultures (McLean et al. 1999;<br />

Chambers et al. 2000) or as ITS-restriction fragment length polymorphism<br />

(RFLP) groups (Midgley et al. 2002, 2004a) and ITS sequences obtained<br />

for representative isolates of each morphological or RFLP group. Comparison<br />

of the ITS sequences with those available in the GenBank nucleotide<br />

database has confirmed that most endophytes isolated from epacrid roots<br />

are ascomycetes or their anamorphs. Indeed, there is only a single published<br />

account of isolation of a fungus from epacrid hair roots that appears,<br />

on the basis of ITS sequence identity, to be a basidiomycete (Midgley<br />

et al. 2004a). This isolate did not, however, form ericoid mycorrhiza in<br />

gnotobiotic culture with Woollsia pungens (see below) and seems most<br />

likely to represent a facultative root inhabitant. Recent analysis of epacrid<br />

endophytes by direct DNA extraction from hair roots has revealed the<br />

presence of other putative basidiomycetes. These, however, appear to be<br />

relatively uncommon and their mycorrhizal status is unclear (D. Bougoure<br />

and J.W.G. Cairney, unpublished data).<br />

ITS sequence data suggest that many of the endophytes are Helotiales<br />

ascomycetes, several of which have affinities with the Hymenoscyphus ericae<br />

aggregate (McLean et al. 1999; Chambers et al. 2000; Sharples et al.<br />

2000; Cairney and Ashford 2002). This aggregate encompasses H. ericae,<br />

Phialophora finlandia and related taxa, many of which form ericoid mycorrhiza<br />

with northern hemisphere Ericaceae (Vrålstad et al. 2002). These isolateshavebeenobtainedfromepacridsatalpine,drysclerophyllforest,sand<br />

mine and coastal heathland sites, indicating that they have a widespread<br />

distribution in a variety of Australian habitats (McLean et al. 1998, 1999;<br />

Midgley et al. 2002). Many of the other endophytes are part of a poorly<br />

defined Helotiales ascomycete group that probably incorporates a number<br />

of taxa and includes many ericoid mycorrhizal and other root-associated<br />

fungi from northern hemisphere Ericaceae (McLean et al. 1999; Chambers<br />

et al. 2000; Sharples et al. 2000; Cairney and Ashford 2002; Berch et al. 2002;<br />

Midgley et al. 2002, 2004a). Other isolates had closest ITS sequence similarity<br />

to Capronia (Chaetothyriales) and Thielavia (Sordariales) (Midgley<br />

et al. 2002, 2004a), and, while some of these are also endophytes of nonericaceae<br />

hosts, similar isolates are known to form ericoid mycorrhiza with<br />

Gaultheria shallon in Canada (Berch et al. 2002). Oidiodendron spp. are<br />

widespread ericoid mycorrhizal fungi of northern hemisphere Ericaceae


250 J.W.G. Cairney<br />

(Perotto et al. 2002; see Chap. 13 by Rice and Currah). Endophytes with<br />

strong ITS sequence similarity to O. maius have recently been isolated, and<br />

may represent common endophytes of certain epacrids (Chambers et al.<br />

2000; D. Bougoure and J.W.G. Cairney, unpublished data).<br />

In addition to the putative ericoid mycorrhizal fungi, many fungi isolated<br />

from surface-sterilised epacrid roots show closest ITS sequence matches to<br />

fungi that are not known to be mycorrhizal fungi. These include isolates<br />

that have closest sequence matches to taxa that are regarded as saprotrophic<br />

or pathogenic, but also with dark septate endophytes (DSE) (Midgley et<br />

al. 2002, 2004a; Davies et al. 2003; D. Bougoure and J.W.G. Cairney, unpublished<br />

data). Isolates with close sequence similarity to Phialocephala<br />

fortinii and other DSE taxa, although occasionally obtained from lower elevation<br />

habitats, have been obtained primarily from epacrids in Australian<br />

sub-alpine and alpine habitats (Davies et al. 2003; Midgley et al. 2004a).<br />

Furthermore, Davies et al. (2003) found that infection of some epacrid<br />

roots by DSE at alpine sites can be more prevalent than ericoid mycorrhizal<br />

infection. Since only limited sampling has so far been undertaken in alpine<br />

habitats, and systematic comparisons of endophyte abundance in different<br />

habitats have yet to be conducted, the ecological significance of these<br />

observations is difficult to assess.<br />

These observations are, however, based on fungi that have been isolated<br />

from epacrid hair roots and assume that all endophytic fungi are isolated<br />

with equal efficiency, or indeed are isolated at all. It is possible, for example,<br />

that fast-growing endophytes may mask slower growing endophytes that<br />

are present in the same root piece, resulting in only the faster-growing taxa<br />

being isolated (Hambleton and Currah 1997). Endophytes that are difficult<br />

or impossible to culture may also be present in epacrid hair roots. This<br />

may be the case in the North American Ericaeae taxon Gaultheria shallon,<br />

for which direct DNA extraction and subsequent cloning of ITS sequences<br />

revealed the presence of a Sebacina-like basidiomycete that was not present<br />

in cultured assemblages of fungi from the same roots (Berch et al. 2002;<br />

Allen et al. 2003). Furthermore, these sequences were obtained from root<br />

segments that appeared to have mycorrhizal infection in epidermal cells,<br />

yet did not yield culturable endophytes. In contrast, where culturable endophyte<br />

assemblages from the Epacris pulchella hair roots were compared<br />

to fungal sequences obtained following direct DNA extraction from roots,<br />

the most-commonly isolated fungi were also commonly represented in<br />

the directly extracted DNA (D. Bougoure and J.W.G. Cairney, unpublished<br />

data). This clearly suggests that in the case of these E. pulchella plants the<br />

most common endophytes were culturable and that the observations from<br />

G. shallon do not serve as a paradigm for Ericaceae in general.


14 Mycorrhizal and Endophytic Fungi of Epacrids (Ericaceae) 251<br />

14.3<br />

Diversity and Spatial Distribution<br />

of Endophyte Taxa in Epacrid Root Systems<br />

Wheremultipleisolationshavebeenmadefromrootsystemsofindividual<br />

epacrids in the field, an assemblage of endophyte taxa has invariably<br />

been recovered (McLean et al. 1999; Chambers et al. 2000). Midgley et al.<br />

(2002, 2004a) undertook intensive isolations from 5.0 mm long root pieces<br />

from throughout the hair root systems of two Woollsia pungens plants and<br />

a Leucopogon parviflorus plant at a sclerophyll forest site in temperate eastern<br />

Australia. Between 99 and 227 isolates were obtained from each plant<br />

and ITS-RFLP analysis suggested that the assemblage from each plant comprised<br />

5 to 17 taxa. Most isolates (76–85%) in each assemblage from a single<br />

plant, however, were of the same ITS-RFLP-type, with a single RFLP-type<br />

found to dominate the root system of the L. parviflorus and a neighbouring<br />

W. pungens plant. Mapping the distribution of the RFLP types according to<br />

their position in the root systems from which they were isolated, demonstrated<br />

that the isolates that dominated the assemblages were widespread<br />

throughout the hair root systems (Midgley et al. 2002, 2004a). These dominant<br />

RFLP-types were shown to form typical ericoid mycorrhiza-like coil<br />

structures in gnotobiotic mycorrhizal infection experiments with W. pungens<br />

as host (Midgley 2003; Midgley et al. 2004a), suggesting that each<br />

root system was dominated by a single putative ericoid mycorrhizal fungal<br />

taxon. Only a few other RFLP-types from each root system formed<br />

mycorrhiza-like coils in W. pungens hair roots, with the remainder failing<br />

to form typical mycorrhizal structures in epidermal cells.<br />

Genetic diversity in populations of the dominant putative ericoid mycorrhizal<br />

fungal taxa isolated from epacrid roots has been investigated using<br />

inter-simple sequence repeat (ISSR) PCR (Midgley et al. 2002, 2004a). These<br />

investigations revealed that, in the case of W. pungens and L. parviflorus<br />

in a dry sclerophyll forest habitat, the population of the dominant taxon<br />

within a root system comprised three to six genotypes. The population<br />

from each root system was, however, dominated (81–96% of isolates) by<br />

a single, spatially widespread genotype of the most abundant putative mycorrhizal<br />

taxon, with the remaining genotypes being relatively uncommon.<br />

In the neighbouring W. pungens and L. parviflorus plants dominated by<br />

the same putative ericoid mycorrhizal fungal taxon, each root system was<br />

dominated by a different genotype of that taxon (Midgey et al. 2004a). The<br />

existence of genotypes that are widespread within root systems indicates<br />

that there is a high probability that isolates from individual root segments<br />

in the same root system will be from the same mycelial individual. This will<br />

complicate investigations of the community ecology of ericoid mycorrhizal


252 J.W.G. Cairney<br />

fungi, particularly where comparisons of relative abundance of taxa between<br />

sites is concerned, and may ultimately inform the spatial scale at<br />

which future sampling should be conducted.<br />

As proposed for populations of the DSE Phialocephala fortinii associated<br />

with roots in a conifer stand, domination of the root system by a single<br />

genotype might result from early establishment and/or greater competitiveness<br />

of the dominant genotype (Grünig et al. 2002). In either case, local<br />

micro-scale edaphic conditions are likely to play a strong selective role in<br />

structuring the endophyte communities and populations. Putative mycorrhizal<br />

fungi from epacrids vary in their responses to water stress (Hutton et<br />

al. 1996b; Chen et al. 2003). Furthermore, the extent of ericoid mycorrhizalike<br />

hair root colonisation varies seasonally for W. pungens in eastern Australian<br />

sclerophyll forests (Kemp et al. 2003). Indeed, in epacrids that are<br />

subject to a Mediterranean-type climate in south-western Australia, few<br />

hair roots survive the summer and apparent ericoid mycorrhizal infection<br />

disappears during the driest months. Infection then increases progressively<br />

with decreasing temperature and increasing rainfall (Hutton et al. 1994).<br />

Although no investigations have so far been conducted, it is thus possible<br />

that the relative abundance of different taxa and/or genotypes within root<br />

systems will be found to show considerable seasonal variation.<br />

The same genotype of a single putative ericoid mycorrhizal fungal taxon<br />

was found to be present in the root systems of the neighbouring W. pungens<br />

and L. parviflorus plants (Midgey et al. 2004a). Interestingly, Liu et<br />

al. (1998) isolated identical genotypes of what, based on the >99% ITS<br />

sequence similarity between the two (Midgley et al. 2004a), appears likely<br />

to be the same taxon identified by Midgley et al. (2004a) from hair roots of<br />

neighbouring W. pungens plants at a different sclerophyll forest site. This<br />

implies that the phenomenon is not uncommon for putative mycorrhizal<br />

fungi of these epacrids, but whether it reflects the presence of single genets<br />

that were continuous between the two plants, or ramets that were confined<br />

to one or other root system remains unresolved. In contrast to ecto- and<br />

arbuscular-mycorrhizal fungi, ericoid mycorrhizal fungi are generally regarded<br />

as producing only limited mycelial growth in soil (Smith and Read<br />

1997). This tenet is, however, based primarily on observations of H. ericae<br />

in mor-humus heathland vegetation communities in the northern hemisphere.<br />

It is possible that some ericoid mycorrhizal fungi of epacrids are<br />

capable of growing further from a host root into soil. Even if this is not<br />

the case, it is possible that the root systems of neighbouring epacrids were<br />

juxtaposed and that hyphal bridges do in fact interconnect epacrid plants.<br />

Such interconnectedness might have significant physiological and ecological<br />

consequences for the symbiotic partners (Robinson and Fitter 1999),<br />

and further investigation in this area is clearly warranted.


14 Mycorrhizal and Endophytic Fungi of Epacrids (Ericaceae) 253<br />

14.4<br />

Mycorrhizal Status of Endophytes from Epacrids<br />

Due to problems of germinating seeds and maintaining epacrid seedlings<br />

under sterile conditions in the laboratory, some of these experiments<br />

have been conducted using northern hemisphere Ericaceae hosts such<br />

as Vaccinium macrocarpon or V. corymbosum (Liu et al. 1998; Davies et<br />

al. 2003). Recent success in sterile propagation of epacrids such as Epacris<br />

impressa by micropropagation and Woollsia pungens from seed, however,<br />

have facilitated the use of epacrids as hosts for mycorrhiza infection testing<br />

(Fig. 14.1a,b) (McLean et al. 1998; Midgley et al. 2004a). ITS sequence data<br />

indicate that those endophytes that have been shown to produce ericoid<br />

mycorrhiza-like infection have affinity with either the Hymenoscyphus ericae<br />

aggregate or two broad groups of Helotiales ascomycetes (equivalent<br />

to “Assemblage A” and “Unknown 2 & possible relatives” in Berch et al.<br />

2002) (McLean et al. 1998; Chambers et al. 2000; Berch et al. 2002; Davies<br />

et al. 2003; Midgley et al. 2004a). Endophytes in the H. ericae aggregate<br />

and both Helotiales groups have also been isolated from, and shown to<br />

form ericoid mycorrhizas with, northern hemisphere Ericaceae (Berch et<br />

al. 2002), suggesting that these groups are important mycorrhizal fungi of<br />

Ericaceae worldwide.<br />

While O. maius and some Thielavia-like and Capronia-like endophytes<br />

from northern hemisphere Ericaceae have been shown to be ericoid<br />

Fig.14.1. a,b Testing for ericoid mycorrhiza formation by endophytes isolated from epacrid<br />

hair roots in gnotobiotic culture. a Woollsia pungens seedling in gnotobiotic culture with<br />

ericoid mycorrhizal fungi. b W. pungens hair root showing dense hyphal coils produced by<br />

an ericoid mycorrhizal fungus in epidermal cells (arrows) Bars a 1cm,b 50 µm


254 J.W.G. Cairney<br />

mycorrhizal fungi (Berch et al. 2002), none of the isolates from epacrids that<br />

have close sequence identity to these fungi formed mycorrhiza-like structures<br />

in gnotobiotic infection experiments (Chambers et al. 2000; Midgley et<br />

al. 2004a). Certain Thielavia-like and Capronia-like isolates from northern<br />

hemisphere Ericaceae hosts also appear to be non-mycorrhizal, suggesting<br />

that further work is required to ascertain the extent of mycorrhiza-forming<br />

abilitiesofisolatesinthesegroups.Theonlybasidiomycetesofarisolated<br />

from an epacrid hair root system failed to form ericoid mycorrhiza-like<br />

structures in W. pungens roots (Midgley et al. 2004a), suggesting that it,<br />

and the basidiomycete hyphae observed in Dracophyllum secundum epidermal<br />

cells by Allen et al. (1989), was probably a saprotroph. Although<br />

they can infect Ericaceae epidermal cells, the coils formed by isolates from<br />

epacrids that have close ITS sequence identity to Phialocephala fortinii<br />

are typical of those formed by DSE rather than ericoid mycorrhizal fungi<br />

(Davies et al. 2003; Midgley et al. 2004a). Interactions between DSE and<br />

their plant hosts remain poorly understood, and it appears that they may,<br />

under different circumstances, have mildly parasitic, neutral, or mildly<br />

beneficial effects on their hosts [Jumpponen 2001; see Chaps. 7 (Sieber and<br />

Grünig) and 15 (Schulz)]. The significance of infection of epacrids by DSE<br />

thus remains unclear.<br />

The broad taxonomic congruence between the fungi that appear to form<br />

ericoid mycorrhizas with epacrids and those that form associations with<br />

Ericaceae from other continents is consistent with the hypothesis that ericoid<br />

mycorrhizal associations evolved in a common ancestral host plant<br />

group (Cullings 1996). Fossil evidence and extant biogeographical patterns<br />

suggest that the association probably arose in Gondwanaland during<br />

the Cretaceous period, and that hosts and associated fungi subsequently<br />

radiated northwards (Cullings 1996; Cairney 2000).<br />

14.5<br />

Saprotrophic Potential of Mycorrhizal Fungi<br />

The presence of propagules of putative ericoid mycorrhizal fungi in sclerophyll<br />

forest soil has been demonstrated by direct DNA extraction from soil<br />

from which roots were removed (Chen and Cairney 2002). Similarly, in the<br />

seasonally drought-affected soils of Western Australian heathlands, ericoid<br />

mycorrhizal inoculum appears to persist through summer in surface soil<br />

above the zone of most hair root growth, again suggesting the presence<br />

of host-free fungal propagules (Hutton et al. 1996a). Further evidence of<br />

the abilities of ericoid mycorrhizal fungi to persist in the absence of their<br />

Ericaceae hosts has recently been obtained for the northern hemisphere<br />

taxon Erica arborea, which became infected by ericoid mycorrhizal fungi


14 Mycorrhizal and Endophytic Fungi of Epacrids (Ericaceae) 255<br />

when planted in a mature Quercus ilex forest that lacked Ericaceae vegetation<br />

(Bergero et al. 2003). Unfortunately the techniques used in these<br />

studies cannot discriminate between actively growing mycelia and inactive<br />

propagules such as asexual spores. However, the activities of the fungi in<br />

axenic culture suggest that, to some extent at least, they are capable of<br />

saprotrophic growth in the absence of an Ericaceae host.<br />

It is well established that, via production of an array of extracellular<br />

enzymes, H. ericae has considerable ability to exist as a free-living saprotroph<br />

(reviewed by Cairney and Burke 1998), and there is evidence that<br />

other mycorrhizal fungi of northern hemisphere Ericaceae have similar<br />

abilities (Piercey et al. 2002; Varma and Bonfante 1994). Relatively little is<br />

known regarding the saprotrophic potential of endophytes from epacrids.<br />

Midgley et al. (2004c), however, have shown that two frequently isolated putative<br />

ericoid mycorrhizal fungal taxa from W. pungens can utilise a range<br />

of compounds as sole carbon sources during growth in axenic culture.<br />

Thus, along with hexoses, these taxa can derive carbon from xylan and<br />

cellulose, suggesting production of 1-4-β-xylanase and β-d-xylosidase,<br />

along with a complete cellulase complex (Midgley et al. 2004c). While<br />

these enzyme activities are doubtless important in the penetration of host<br />

epidermal cell walls during the establishment of mycorrhizal symbiosis,<br />

they probably also facilitate a degree of saprotrophic growth and may<br />

be functionally important in the process of symbiotic nutrient acquisition.<br />

14.6<br />

Symbiotic Functioning of Mycorrhizal Fungi<br />

The mutualistic nature of ericoid mycorrhizal infection of epacrids has not<br />

yet been confirmed by gnotobiotic culture nutrient transfer experiments<br />

(see McLean et al. 1998; Anthony et al. 2000; Cairney and Ashford 2002).<br />

Nonetheless, the putative ericoid mycorrhizal fungi can utilise a range of<br />

amino acids and simple proteins as sole nitrogen sources, along with inositol<br />

hexaphosphate and DNA as sole sources of phosphorus (Chen et al.<br />

1999; Whittaker and Cairney 2001; Midgley et al. 2004b). Their abilities to<br />

access nitrogen and phosphorus thus appear to be on a par with H. ericae<br />

and other ericoid mycorrhizal fungi from northern hemisphere Ericaceae,<br />

which are known to acquire nitrogen and phosphorus from these substrates<br />

and effect transfer of the elements to plant hosts (see Smith and<br />

Read 1997; Xiao and Berch 1999). Given the relatedness of the putative<br />

ericoid mycorrhizal fungi of epacrids to these known mycorrhizal fungi,<br />

the similarities in the infections formed in host epidermal cells and their<br />

abilities to utilise organic nitrogen and phosphorus sources, however, it


256 J.W.G. Cairney<br />

seems probable that they function in a manner similar to their northern<br />

hemisphere counterparts.<br />

In addition to utilising organic nitrogen substrates, Midgley et al. (2004b)<br />

found that, as is the case for H. ericae, putative mycorrhizal fungi from<br />

epacrids are efficient users of nitrate. This contrasts with ectomycorrhizal<br />

fungi from the same eastern Australian sclerophyll forest habitats, which<br />

have limited abilities to utilise nitrate (Anderson et al. 1999; Sawyer et<br />

al. 2003). Although the soils in such forests are likely to be ammonifying<br />

(Connell et al. 1995), the ability to utilise nitrate may be advantageous to<br />

the mycorrhizal fungi and their epacrid hosts following fire, when nitrate<br />

may be relatively abundant (Stewart et al. 1993). Many epacrids are killed,<br />

but regenerate rapidly from seed, following fire, while others may survive<br />

fire by resprouting (Bell et al. 1996). There is also evidence that epacrid<br />

abundance increases following low intensity fires (Morrison 2002) and<br />

that putative ericoid mycorrhizal fungi can be present in some sclerophyll<br />

forest soils within a few days of a low intensity fire (Chen and Cairney 2002).<br />

Since non-mycorrhizal epacrids have only limited abilities to utilise NO − 3<br />

for growth (Stewart et al. 1993), these observations suggest that the fungi<br />

may be of particular importance to the success of epacrids in fire-prone<br />

vegetation communities.<br />

It has been postulated that different mycorrhizal fungal taxa or genotypes<br />

might differentially access nutrient sources in soil, and that this might<br />

be important in increasing overall nitrogen and/or phosphorus uptake by<br />

their plant hosts (Cairney et al. 2000; Koide 2000). Midgley et al. (2004b)<br />

investigated the relative abilities of six genotypes of an H. ericae-like putative<br />

mycorrhizal fungus from W. pungens and six isolates of an unknown<br />

Helotiales putative mycorrhizal fungal taxon from W. pungens and L. parviflorus<br />

to utilise a range of simple inorganic and organic nitrogen and<br />

phosphorus sources for growth. All isolates were found to utilise inorganic<br />

forms of the elements, acidic, neutral and basic amino acids, along with<br />

protein, phosphomonoesterase and phosphodiesterase. Although some intraspecific<br />

variation was observed on all substrates, one taxon produced<br />

significantly more biomass on most substrates than the other. For all substrates,<br />

however, the difference between the two taxa was considerably less<br />

than an order of magnitude, leading Midgley et al. (2004b) to conclude<br />

that both taxa would confer broadly similar nutritional benefits to their<br />

hosts. This, and the fact that screening of single isolates of a range of putative<br />

ericoid mycorrhizal fungal taxa from W. pungens suggests that all<br />

are broadly similar in their abilities to utilise organic nitrogen and phosphorus<br />

substrates (Chen et al. 1999; Whittaker and Cairney 2001), might<br />

appear to contradict the proposal that different isolates/genotypes vary in<br />

their abilities to enhance host nutrition. This, however, may not be the<br />

case. The efficiency with which the various fungi transfer nitrogen and/or


14 Mycorrhizal and Endophytic Fungi of Epacrids (Ericaceae) 257<br />

phosphorus from the substrates to the host plant remains to be tested, and<br />

different taxa/genotypes may differ considerably in this respect.<br />

14.7<br />

Conclusions<br />

Although research to date has been confined to a small number of epacrid<br />

taxa from a limited range of habitats, it appears that most endophytes that<br />

have been isolated from epacrid hair roots are probably ericoid mycorrhizal<br />

fungi. An array of mainly Helotiales ascomycetes forms putative ericoid<br />

mycorrhizal associations with epacrids, but root systems of individual<br />

plants in the field are dominated by a small number of taxa, and populations<br />

of these dominated by single genotypes. Studies of mycorrhizal fungal<br />

diversity in natural habitats are at an early stage and in the future will<br />

need to take into account the spatial distribution of mycelial individuals<br />

if meaningful comparisons of communities in different habitats are to be<br />

made. Functional aspects of the symbiotic relationships between putative<br />

ericoid mycorrhizal fungi and their epacrid hosts remain to be investigated,<br />

however it is likely that these will follow the established paradigm for ericoid<br />

mycorrhizal associations of northern hemisphere Ericaecae.<br />

Acknowledgements. I thank D.J. Midgley for his permission to use Fig. 14.1b.<br />

<strong>References</strong><br />

Allaway WG (1996) Biology of the Epacridaceae. Ann Bot 77:290–291<br />

Allen TR, Millar T, Berch SM, Berbee ML (2003) Culturing and direct DNA extraction find<br />

different fungi from the same ericoid mycorrhizal roots. New Phytol 160:255–272<br />

Allen WK, Allaway WG, Cox GC, Valder PG (1989) Ultrastructure of mycorrhizas of Dracophyllum<br />

secundum R. Br. (Ericales: Epacridaceae). Aust J Plant Physiol 16:147–153<br />

Anderson IC, Chambers SM, Cairney JWG (1999) Intra- and interspecific variation in patterns<br />

of organic and inorganic nitrogen utilisation by three Australian Pisolithus species.<br />

Mycol Res 103:1579–1587<br />

Anthony J, McLean CB, Lawrie AC (2000) In vitro propagation of Epacris impressa<br />

(Epacridaceae) and the effects of clonal material. Aust J Bot 48:215–221<br />

Bell TL, Pate JS, Dixon KW (1996) Relationships between fire response, morphology, root<br />

anatomy and starch distribution in south-western Australian Epacridaceae. Ann Bot<br />

77:357–364<br />

Bellgard SE (1991) Mycorrhizal associations of plant species in Hawkesbury sandstone<br />

vegetation. Aust J Bot 39:357–364<br />

Berch SM, Allen TR, Berbee ML (2002) Molecular detection, community structure and<br />

phylogeny of ericoid mycorrhizal fungi. Plant Soil 244:55–66<br />

Bergero R, Girlanda M, Bello F, Luppi AM, Perotto S (2003) Soil persistence and biodiversity<br />

of ericoid mycorrhizal fungi in the absence of the host plant in a Mediterranean<br />

ecosystem. Mycorrhiza 13:69–75


258 J.W.G. Cairney<br />

Briggs CL, Ashford AE (2001) Structure and composition of the thick wall in hair root<br />

epidermal cells of Woollsia pungens. New Phytol 149:219–232<br />

Cairney JWG (2000) Evolution of mycorrhiza systems. Naturwissenschaften 87:467–475<br />

Cairney JWG, Ashford AE (2002) Biology of mycorrhizal associations of epacrids (Ericaceae).<br />

New Phytol 154:305–326<br />

Cairney JWG, Burke RM (1998) Extracellular enzyme activities of the ericoid mycorrhizal<br />

endophyte Hymenoscyphus ericae (Read) Korf & Kernan: their likely roles in decomposition<br />

of dead plant material in soil. Plant Soil 205:181–192<br />

Cairney JWG, Sawyer NA, Sharples JM, Meharg AA (2000) Intraspecific variation in nitrogen<br />

source utilisation by isolates of the ericoid mycorrhizal fungus Hymenoscyphus ericae<br />

(Read) Korf & Kernan. Soil Biol Biochem 32:1319–1322<br />

Chambers SM, Liu G, Cairney JWG (2000) ITS rDNA sequence comparison of ericoid<br />

mycorrhizal endophytes from Woollsia pungens. Mycol Res 104:168–174<br />

Chen A, Chambers SM, Cairney JWG (1999) Utilisation of organic nitrogen and phosphorus<br />

sources by mycorrhizal endophytes of Woollsia pungens (Cav.) F. Muell. (Epacridaceae).<br />

Mycorrhiza 8:181–187<br />

Chen DM, Cairney JWG (2002) Investigation of the influence of prescribed burning on ITS<br />

profiles of ectomycorrhizal and other soil fungi at three Australian sclerophyll forest<br />

sites. Mycol Res 106:532–540<br />

Chen DM, Khalili K, Cairney JWG (2003) Influence of water stress on biomass production<br />

by isolates of an ericoid mycorrhizal endophyte of Woollsia pungens and Epacris<br />

microphylla (Ericaceae). Mycorrhiza 13:173–176<br />

Connell MJ, Raison RJ, Khanna PK (1995) Nitrogen mineralisation in relation to site history<br />

and soil properties for a range of Australian forest soils. Biol Fert Soil 20:213–220<br />

Copeland HF (1954) Observations on certain Epacridaceae. Am J Bot 41:215–222<br />

Cullings KW (1996) Single phylogenetic origin of ericoid mycorrhizae within the Ericaceae.<br />

Can J Bot 74:1896–1909<br />

Davies PW, Mclean CB, Bell TL (2003) Root survey and isolation of fungi from alpine<br />

epacrids (Ericaceae). Aust Mycol 22:4–10<br />

Grünig CR, Sieber TN, Rogers SO, Holdernrieder O (2002) Spatial distribution of dark<br />

septate endophytes in a confined forest plot. Mycol Res 106:832–840<br />

Hambleton S, Currah RS (1997) Fungal endophytes from the roots of alpine and boreal<br />

Ericaceae. Can J Bot 75:1570–1581<br />

Hutton BJ, Dixon KW, Sivasithamparam K (1994) Ericoid endophytes of Western Australian<br />

heaths (Epacridaceae). New Phytol 127:557–566<br />

Hutton BJ, Dixon KW, Sivasithamparam K, Pate JS (1996a) Inoculum potential of ericoid<br />

endophytes of Western Australian heaths (Epacridaceae). New Phytol 134:665–672<br />

Hutton BJ, Sivasithamparam K, Dixon KW, Pate JS (1996b) Pectic zymograms and water<br />

stress tolerance of endophytic fungi isolated from Western Australian heaths<br />

(Epacridaceae). Ann Bot 77:399–404<br />

Jumpponen A (2001) Dark septate endophytes – are they mycorrhizal? Mycorrhiza<br />

11:207–211<br />

Keighery GJ (1996) Phytogeography, biology and conservation of Western Australian<br />

Epacridaceae. Ann Bot 77:347–355<br />

Kemp E, Adam P, Ashford AE (2003) Seasonal changes in hair roots and mycorrhizal<br />

colonization in Woollsia pungens (Cav.) F. Muell. (Epacridaceae). Plant Soil 250:241–248<br />

Khan AG (1978) Vesicular arbuscular mycorrhizas in plants colonising black wastes from<br />

bituminous coal mining in the Illawarra region of New South Wales. New Phytol 81:53–63<br />

Koide RT (2000) Functional complementarity in the arbuscular mycorrhizal symbiosis. New<br />

Phytol 147:233–235


14 Mycorrhizal and Endophytic Fungi of Epacrids (Ericaceae) 259<br />

Kron KA, Judd WS, Stevens PF, Crayn DM, Anderberg AA, Gadek PA, Quinn CJ, Luteyn JL<br />

(2002) Phylogenetic classification of Ericaceae: molecular and morphological evidence.<br />

Bot Rev 68:335–423<br />

Liu G, Chambers SM, Cairney JWG (1998) Molecular diversity of ericoid mycorrhizal endophytes<br />

isolated from Woollsia pungens. New Phytol 140:145–153<br />

McGee P (1986) Mycorrhizal associations of plant species in a semiarid community. Aust<br />

J Bot 34:585–593<br />

McLean C, Lawrie AC (1996) Patterns of root colonization in epacridaceous plants collected<br />

from different sites. Ann Bot 77:405–411<br />

McLean CB, Anthony J, Collins RA, Steinke E, Lawrie AC (1998) First synthesis of ericoid<br />

mycorrhizas in the Epacridaceae under axenic conditions. New Phytol 139:589–593<br />

McLean CB, Cunnington JH, Lawrie AC (1999) Molecular diversity within and between<br />

ericoid endophytes from the Ericaceae and Epacridaceae. New Phytol 144:351–358<br />

Midgley DJ (2003) Diversity and distribution of mycorrhizal and root-associated fungal<br />

endophytes in Leucopogon parviflorus and Woollsia pungens (Ericaceae). PhD Thesis,<br />

University of Western Sydney, Australia<br />

Midgley DJ, Chambers SM, Cairney JWG (2002) Spatial distribution of fungal endophyte<br />

genotypes in a Woollsia pungens (Ericaceae) root system. Aust J Bot 50:559–565<br />

Midgley DJ, Chambers SM, Cairney JWG (2004a) Distribution of ericoid mycorrhizal endophytes<br />

and root-associated fungi in neighbouring Ericaceae plants in the field. Plant<br />

Soil 259:137–151<br />

Midgley DJ, Chambers SM Cairney JWG (2004b) Inorganic and organic substrates as sources<br />

of nitrogen and phosphorus for multiple genotypes of two ericoid mycorrhizal fungal<br />

taxa from Woollsia pungens Cav. (Muell.) and Leucopogon parviflorus (Andr.) Lindl.<br />

(Ericaceae). Aust J Bot 52:63–71<br />

Midgley DJ, Chambers SM, Cairney JWG (2004c) Utilisation of carbon substretae by multiple<br />

genotypes of ericoid mycorrhizal fungal endophytes from eastern Australian Ericaceae.<br />

Mycorrhiza 14:245–251<br />

Morrison DA (2002) effects of fire intensity on plant species composition of sandstone<br />

communities in the Sydney region. Aust Ecol 27:433–441<br />

Perotto S, Girlanda M, Martino E (2002) Ericoid mycorrhizal fungi: some new perspectives<br />

on old acquaintances. Plant Soil 244:41–53<br />

Piercey MM, Thormann MN, Currah RS (2002) Saprobic characteristics of three fungal<br />

taxa from ericalean roots and their association with the roots of Rhododendron groenlandicum<br />

and Picea mariana in culture. Mycorrhiza 12:175–180<br />

Powell JM, Crayn DM, Gadek PA, Quinn CJ, Morrison DA, Chapman AR (1996) A reassessment<br />

of relationships within Epacridaceae. Ann Bot 77:305–315<br />

Read DJ (1996) The structure and function of the ericoid mycorrhizal root. Ann Bot<br />

77:365–374<br />

Reed ML (1989) Ericoid mycorrhizas of Styphelieae: intensity of infection and nutrition of<br />

the symbionts. Aust J Plant Physiol 16:155–160<br />

Reed ML (1996) Diversity of mycorrhizal fungi in the roots of epacrids. In: Hopper SD,<br />

Chappill JA, Harvey MS, George AS (eds) Gondwanan heritage: past, present and<br />

future of the Western Australian biota. Surrey Beattie, Chipping Norton, Australia,<br />

pp 309–313<br />

Robinson D, Fitter A (1999) The magnitude and control of carbon transfer between plants<br />

linked by a common network. J Exp Bot 50:9–13<br />

Sawyer NA, Chambers SM, Cairney JWG (2003) Utilisation of inorganic and organic nitrogen<br />

sources by Amanita species native to temperate eastern Australia. Mycol Res<br />

107:413–420


260 J.W.G. Cairney<br />

Sharples JM, Chambers SM, Meharg AA, Cairney JWG (2000) Genetic diversity of rootassociated<br />

fungal endophytes from Calluna vulgaris at contrasting field sites. New<br />

Phytol 148:153–162<br />

Smith SE, Read DJ (1997) Mycorrhizal symbiosis. Academic Press, London<br />

Steinke E, Williams PG, Ashford AE (1996) The structure and fungal associates of mycorrhizas<br />

in Leucopogon parviflorus (Andr.) Lindl. Ann Bot 77:413–419<br />

Stewart GR, Pate JS, Unkovich M (1993) Characteristics of inorganic nitrogen assimilation<br />

of plants in fire-prone Mediterranean type vegetation. Plant Cell Environ 16:351–363<br />

Varma A, Bonfante P (1994) Utilisation of cell-wall related carbohydrates by ericoid mycorrhizal<br />

endophytes. Symbiosis 16:301–313<br />

Vrålstad T, Myhre E, Schumacher T (2002) Molecular diversity and phylogenetic affinities<br />

of symbiotic root-associated ascomycetes of the Helotiales in burnt and metal polluted<br />

habitats. New Phytol 155:131–148<br />

Whittaker SP, Cairney JWG (2001) Influence of amino acids on biomass production by<br />

ericoid mycorrhizal endophytes from Woollsia pungens (Epacridaceae). Mycol Res<br />

105:105–111<br />

Xiao G, Berch SM (1999) Organic nitrogen use by salal ericoid mycorrhizal fungi from northern<br />

Vancouver Island and impacts on growth in vitro of Gaultheria shallon. Mycorrhiza<br />

9:145–149


15<br />

15.1<br />

Introduction<br />

Mutualistic Interactions<br />

with Fungal Root Endophytes<br />

Barbara Schulz<br />

With few exceptions, colonisation of a plant host is beneficial for the fungus,<br />

assuring it a supply of nutrients and shelter from most abiotic stressors.<br />

Previously, within plant roots only the symbioses of mycorrhizal fungi<br />

were considered to be mutualistic. Recently, it has been recognised that<br />

many other fungi, and in particular endophytic fungi, can participate in<br />

mutualistic root symbioses (e.g. Sieber 2002; Brundrett 2002; Schulz and<br />

Boyle 2005).<br />

There are various potential benefits for the host in mutualistic interactions<br />

with endophytic fungi, for example induction of defence metabolites<br />

potentially active against pathogens (Schulz et al. 1999; Mucciarelli et al.<br />

2003; Arnold and Herre 2003), endophytic secretion of phytohormones<br />

(Holland 1997; Rey et al. 2001; Römmert et al. 2002; Tudzynski and Sharon<br />

2002), mobilisation of nutrients for the host from the rhizosphere (Jumpponen<br />

et al. 1998; Caldwell et al. 2000; Usuki et al. 2002) and/or an alteration<br />

of host metabolism (Jallow et al. 2004). Colonisation by fungal root endophytes<br />

may lead to induced disease resistance (Picard et al. 2000; Benhamou<br />

and Garand 2001), improved growth of the host (Kimura et al. 1992; Jumpponen<br />

2001; Mucciarelli et al. 2002; Ernst et al. 2003), abiotic stress tolerance<br />

(Barrow and Aaltonen 2001; Redman et al. 2002; Barrow 2003) or protection<br />

from pathogenic competitors and insect predators of the host through<br />

synthesis of antagonistic fungal secondary metabolites (Schulz et al. 1995;<br />

Hallmann and Sikora 1996; Schulz et al. 2002; Miller et al. 2002; Selosse et<br />

al. 2004; see Chap. 8 by Bacon and Yates). This chapter reviews results dealing<br />

with mutualistic interactions between non-mycorrhizal root-colonising<br />

endophytic fungi with their plant hosts. When reading this chapter it is important<br />

to bear in mind that within the realms of our present knowledge,<br />

far from all of the interactions with fungal root endophytes are beneficial<br />

for both partners.<br />

Barbara Schulz: Technical University of Braunschweig, Institute of Microbiology, Spielmannstraße<br />

7, 38106 Braunschweig, Germany, E-mail: b.schulz@tu-bs.de<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


262 B. Schulz<br />

15.2<br />

Colonisation and Histology<br />

To the extent that histological studies have been undertaken, growth of<br />

fungal endophytes in roots is usually extensive and can be inter- and/or<br />

intra-cellular (Boyle et al. 2001; Sieber 2002; Schulz and Boyle 2005). Most of<br />

the histological studies have dealt with endophytic colonisation by Fusarium<br />

and dark septate endophytes (DSE; Jumpponen and Trappe 1998; Sieber<br />

2002).<br />

Histological studies of colonisation by DSE have focussed on Phialocephala<br />

fortinii, but also on other DSE including Phialophora spp., other<br />

Phialocephala spp., Chloridium paucisporum, Heteroconium chaetospira<br />

and Leptodontidium orchidicola. The hyphae of DSE vary from thin and<br />

hyaline to melanised, sometimes forming black microsclerotia. The hyaline<br />

hyphae grow both intercellularly and within the vascular cylinder and,<br />

upon favourable nutrient status of the host, contain lipid vacuoles (Barrow<br />

and Aaltonen 2001; Barrow 2003). Barrow (2003) suggests that polymorphic<br />

fungal structures, which are found primarily within the sieve elements and<br />

accumulate massive quantities of lipids, are protoplasts. Not only DSE, but<br />

also other endophytes, e.g. Cryptosporiopsis sp., colonise the vascular cylinder<br />

(Schulz and Boyle 2005). The hyphae of root endophytes often develop<br />

intracellular coils, e.g. H. chaetospira and Oidiodendron maius (Usuki and<br />

Narisawa 2005), as does the basidiomycete, Piriformospora indica (Varma<br />

et al. 2000). These presumably increase the absorption of assimilates or the<br />

exchange of nutrients.<br />

In axenic culture, the DSE Phialocephala colonised the surface and the<br />

inner cortex of the roots of seedlings of Norway spruce and Scotch pine<br />

(Sieber 2002) and Larix decidua (Schulz et al. 1999; A.K. Römmert unpublished)<br />

with mats of mycelia growing within the roots both inter- and<br />

intra-cellularly. From roots of Norway spruce and Scotch pine collected in<br />

the field, the density of mycelial and intracortical development was much<br />

lower (Sieber 2002), suggesting that, under natural conditions, plant defence<br />

limited the density of colonisation. However, in arid habitats, Barrow<br />

(2003) found a highly branched extraradical hyphal network of hyaline<br />

vacuolated hyphae in a mucilaginous gel covering roots of Bouteloua.The<br />

polysaccharide mucilage, which was apparently produced by the fungi,<br />

stores moisture under arid conditions (see Sect. 15.6).<br />

Depending on the host being colonised, some DSE can develop mycorrhizal<br />

structures. For example, in the roots of at least 19 plant species,<br />

the dematiaceous hyphomycete H. chaetospira grew intra- or intercellularly<br />

within the cortical cells (Narisawa et al. 1998; Usuki et al. 2002; Ohki<br />

et al. 2002). However, within the roots of Rhododendron obtusum var.<br />

kaempferi, it developed typical ericoid mycorrhizas (Usuki and Narisawa


15 Mutualistic Interactions with Fungal Root Endophytes 263<br />

2005). P. fortinii has even been found to form a Hartig net and a thin patchy<br />

mantle, considered the anatomical hallmarks of ectomycorrhizae, both in<br />

axenic culture of seedlings of Salix glauca (Fernando and Currah 1996) and<br />

Betula platyphylla (Hashimoto and Hyakumachi 2001), with the roots of<br />

some nursery stocks of Pinus banksiana, Pinus contorta and Pinus glauca<br />

(Danielson and Visser 1990) and with the roots of Populus tremula x Populus<br />

tremuloides grown in an experimental field plot (Kaldorf et al. 2004).<br />

ThisisalsothecasefortheDSEPhialophora finlandia,whichformedectendomycorrhizal<br />

associations with the roots of Pinus resinosa (Lobuglio and<br />

Wilcox 1988) and ectomycorrhizal associations with the roots of several<br />

trees (Vralstad et al. 2002).<br />

The growth modus of avirulent strains of Fusarium spp. is variable,<br />

but often differs from that of virulent strains. Bacon and Hinton (1996)<br />

found that only pathogenic strains of F. moniliforme colonised maize both<br />

inter- and intracellularly; growth of the avirulent isolate was intercellular.<br />

However, other avirulent strains of Fusarium spp. colonised the roots of<br />

barley (Boyle et al. 2001) and pea (Benhamou and Garand 2001) both<br />

inter- and intra-cellularly. Whereas active growth of the avirulent strain<br />

of Fusarium oxysporum was restricted to the root surface, epidermis and<br />

outer cortex of the pea roots, the pathogenic strain of F. oxysporum rapidly<br />

colonised epidermis, cortex, endodermis and paratracheal parenchyma<br />

cells (Benhamou and Garand 2001).<br />

The colonisation modus of Stagonospora spp. in Phragmites australis is<br />

exemplary for fungi that initially colonise the roots, but then apparently<br />

grow systemically within the entire plant following vertical transmission<br />

(Ernst et al. 2003). Similarly, Sieber et al. (1988) isolated Stagonospora nodorum<br />

as an endophyte from both roots and shoots of wheat. Another fungus<br />

that colonises both roots and shoots is a non-sporulating endophyte of<br />

Mentha piperita. Theendophyteformedanexternalenvelopingmycelial<br />

sheet around the root with only little penetration of the root cortex (Mucciarelli<br />

et al. 2003), but was detected in the parenchymatic cells of mature<br />

leaves, especially during senescence (Mucciarelli et al. 2002).<br />

The histology of fungal root colonisations in mutualistic interactions is<br />

extremely variable, in part because any one species can exhibit extreme phenotypic<br />

plasticity. Colonisation can be inter- and/or intra-cellular, limited<br />

to the roots or apparently systemic within the entire plant. The morphologies<br />

vary from apparent protoplasts to very fine undifferentiated hyaline<br />

hyphae to highly differentiated ectomycorrhizas. So, it seems that neither<br />

fungal morphology nor mode of colonisation is decisive for determining<br />

whether or not an interaction is mutualistic. But perhaps the quantity of<br />

colonisation is a decisive factor, since extensive colonisation is common to<br />

all of the investigated mutualistic interactions.


264 B. Schulz<br />

15.3<br />

Secondary Metabolites<br />

A strikingly high proportion of endophytic fungi (80%) produce biologically<br />

active metabolites in vitro in tests for antibacterial, fungicidal and<br />

herbicidal activities (Schulz et al. 2002). Most of the investigations dealing<br />

with the synthesis of endophytic metabolites active against phytopathogens<br />

(Schulz et al. 1995; Tan and Zou 2001; Schulz et al. 2002; Fang-ting et al.<br />

2004) and against predators (Azevedo et al. 2000) have not specified from<br />

whichorgantheendophyteswereisolated.<br />

Fungal secondary metabolites may play a role within the host, and/or<br />

have an ecological significance. As hypothesised by Demain (1980): “If<br />

a fungus can produce metabolites in vitro, they must also have a function<br />

in nature” Fungi would not retain the multienzyme reaction sequences required<br />

for the synthesis of secondary metabolites without some beneficial<br />

effect for survival. As virulence factors, we have hypothesised that secondary<br />

metabolites are involved in maintaining a balance of antagonisms<br />

in the interaction with the host (Schulz et al. 1999; Schulz and Boyle 2005;<br />

see Chap. 1 by Schulz and Boyle). However, they may also have antagonistic<br />

functions.<br />

Some endophytes, for example Fusarium spp., that colonise both the<br />

shoots and roots of numerous hosts, synthesise a number of toxins including<br />

beauvericin, a cyclic hexadepsipeptide with insecticidal properties<br />

(see Chap. 8 by Bacon and Yates; Kuldau and Yates 2000; Miller 2001).<br />

Beauvericins are produced by at least 12 Fusarium species, and protect<br />

infected plants against herbivoric insects (see Chap. 8 by Bacon and Yates).<br />

Fungal root endophytes also synthesise metabolites toxic to the nematode<br />

Meloidogyne incognita. For example, both phomalactone, synthesised<br />

by Verticillium chlamydosporium (Khambay et al. 2000), and metabolites<br />

present in the culture filtrate of a non-pathogenic F. oxysporum reduced<br />

mobility of the nematode within 10 min of exposure (Hallmann and Sikora<br />

1996).<br />

Many root endophytic fungi produce antimicrobial metabolites (Schulz<br />

et al. 2002), e.g. the antibacterial and antifungal metabolite(s) produced in<br />

vitro by Cryptosporiopsis sp. from Larix decidua (Schulz et al. 1995). These<br />

included the antifungal metabolite mycorrhizin, which in situ could protect<br />

the host from phytopathogenic fungi (Schulz et al. 1995). Hallmann and<br />

Sikora (1996) found that the culture filtrate of F. oxysporum strain 162 was<br />

not only toxic to M. incognita, but was also antifungal, inhibiting soil-borne<br />

plant pathogens.<br />

Bultman and Murphy (2000) suggested that endophytic Neotyphodium<br />

spp. are stimulated to increase production of mycotoxins in shoots of<br />

grasses after damage to the host has occurred, the adaptive significance


15 Mutualistic Interactions with Fungal Root Endophytes 265<br />

being apparent. A similar effect could occur when roots colonised by endophytic<br />

fungi are injured. Fungal secondary metabolites may also play roles<br />

in signalling and growth enhancement (see Sect. 15.4).<br />

In conclusion, since most of the fungal root isolates can produce biologically<br />

active secondary metabolites in vitro, it seems probable that they are<br />

also produced in planta. They can be antagonistic against plant predators<br />

and microbial antagonists, in both cases suppressing disease. But they may<br />

also play roles within the host, e.g. in maintaining a balance of antagonism<br />

between endophyte and host.<br />

15.4<br />

Growth Enhancement<br />

The non-mycorrhizal fungal root colonisers that have been reported to<br />

improve growth of their hosts have various growth modi and belong<br />

to diverse genera, including Chaetomium, Cladorrhinum, Cryptosporiopsis,<br />

Fusarium, Heteroconium, Oidiodendron, Phialocephala, Piriformospora<br />

and Stagonospora.<br />

Stagonospora spp., and a non-sporulating endophyte of Mentha piperita,<br />

are exemplary for endophytes that grow systemically in roots and shoots of<br />

their hosts (see Sect. 15.2). When isolates of three species of Stagonospora<br />

were reinoculated into axenic host seedlings of Phragmites australis, all<br />

increased growth significantly (Ernst et al. 2003). The authors note that<br />

this is only the third reported case, besides those dealing with Neotyphodium,<br />

in which a seed-transmitted fungus enhanced the biomass of its<br />

host. Mucciarelli et al. (2002, 2003) speculated that improved growth of<br />

M. piperita might be due to a better nutrient supply or, alternatively, to<br />

the synthesis of plant growth hormones by a non-sporulating endophytic<br />

fungus.<br />

In contrast, colonisation in most of the studied interactions is limited<br />

totheroots,e.g.thatbyPhoma in Vulpia ciliata spp. ambigua, whichincreased<br />

shoot biomass, root lengths, and tiller numbers of the host (Newsham<br />

1994), and that of barley by Chaetomium or Chaetomium globosum,<br />

which increased root fresh weight (Vilich et al. 1998). Gasoni and Stegman<br />

de Gurfinkel (1997) suggested that increased phosphorous uptake, as observed<br />

in cotton roots colonised by Cladorrhinum foecumdissimum, was<br />

responsible for promoting growth of the host. Similarly, results obtained<br />

by Jumpponen and Trappe (1998) led Jumpponen (1999) to the conclusion<br />

that growth enhancement by the DSE may be due to improved phosphorous<br />

and nitrogen uptake (Jumpponen et al. 1998) or, in a closed system,<br />

to increased availability of carbohydrates and/or CO2, both resulting from<br />

fungal metabolism (Jumpponen and Trappe 1998).


266 B. Schulz<br />

It is also possible that growth enhancement is due to indirect acquisition<br />

of nutrients obtained saprophytically from the endophyte from the<br />

rhizosphere. Caldwell et al. (2000) suggested that due to their hydrolytic<br />

capabilities, DSE are able to grow both biotrophically and saprophytically.<br />

Of note: they found no evidence for lignolytic enzymes. They hypothesised<br />

that the fungi may access litter and detrital carbon, nitrogen and phosphorous,<br />

making this available to the host. On the basis of finding a continuous<br />

hyphal network extending from the rhizosphere to the vascular cylinder,<br />

Barrow (2003) also suggested that there is potential for bidirectional carbon<br />

transport. The situation could be similar to that hypothesised for Oidiodendron<br />

maius, which not only develops ericoid mycorrhizas within the<br />

host, but also can grow saprophytically in peat (Rice and Currah 2002,<br />

see Chap. 13 by Rice and Currah) and perhaps supplies its host with organic<br />

nutrients from the rhizosphere. Inter-plant connections may also be<br />

involved in growth enhancement, as suggested by Sen et al. (1999), who<br />

found evidence that a common population of Ceratobasidium cornigerum<br />

occupies roots of orchid and pine. DSE may even replace arbuscular mycorrhiza<br />

(AM) and ectomycorrhizal fungi at sites with extreme environmental<br />

conditions (Sieber 2002; see Chap. 7 by Sieber and Grünig). Co-inoculation<br />

of tomato with a non-pathogenic strain of Fusarium oxysporum and the<br />

AM fungus Glomus coronatum resulted in better mycorrhization, but did<br />

not improve growth more than mono-inoculation (Diedhiou et al. 2003).<br />

Colonisation by the DSE Phialocephala does not always exert positive effects<br />

on the host: in axenic culture it significantly retarded growth of Betula<br />

platyphylla var. japonica seedlings. Nevertheless, P. fortinii formed typical<br />

ectomycorrhiza with B. platyphylla, underlining the plasticity of the interaction<br />

(Hashimoto and Hyakumachi 2001). Whether or not colonisation by<br />

DSE improves growth of the host depends on environmental conditions and<br />

on the metabolic status of the host. Usuki et al. (2002) found that variance<br />

of the growth medium affected the mode of colonisation by Heteroconium<br />

chaetospira in seedlings of Chinese cabbage. Colonisation and growth enhancement<br />

were best, and intracellular, in peat moss amended with 0.1%<br />

glucose. At higher glucose concentrations both frequency of colonisation<br />

and plant growth decreased, and fungal colonisation was restricted mostly<br />

to the intercellular regions of epidermal and cortical cells, leading to the<br />

formation of microslerotia.<br />

Not only colonisation, but also culture extracts of the endophyte can<br />

improve growth. Colonisation of the roots of seedlings of Larix decidua by<br />

the endophytes P. fortinii and Cryptosporiopsis sp., both of which had been<br />

isolated from roots of the host, significantly improved lengths (Schulz et<br />

al. 2002) and dry weights (Römmert et al. 2002) of both roots and shoots<br />

(Fig. 15.1). Additionally, disease symptoms resulting from the stress of axenic<br />

culture decreased (Schulz et al. 1999; Römmert et al. 2002). Similarly,


15 Mutualistic Interactions with Fungal Root Endophytes 267<br />

Fig.15.1. Influence of colonisation on dry weights of shoots (a) and roots (b) of Larix<br />

decidua, cultivated for 3 months axenically in a synthetic medium in expanded clay and<br />

inoculated at day 0 with either an endophyte, Cryptosporiopsis sp. (Cs) or Phialocephala<br />

fortinii (Pf), or a pathogen, Heterobasidion annosum (Ha). n = 29−55, ∗P


268 B. Schulz<br />

Table 15.1. Addition of methanolic culture extract of Phialocephala fortinii grown in MMNA<br />

liquid medium (Schulz et al. 1999) for 14 days to surface-sterilised seedlings of Larix decidua<br />

(approx. 3 months old) and Triticum aestivum (approx. 1 week old) on filter paper, previously<br />

germinated on biomalt agar medium (5% w/v). Evaluation of L. decidua after 28 days, of<br />

T. aestivum after 7 days of incubation. Methanolic extracts of the sterile culture medium<br />

were used as the controls<br />

Average shoot length (cm) Average dry weight of<br />

root+shoot (mg)<br />

L. decidua<br />

+Cultureextract(n = 8) 1.0* 44.0*<br />

Control (n = 10) 0.08 16.9<br />

T. aestivum L.<br />

+Cultureextract(n = 20) 0.84 1.6<br />

Control (n = 20) 0.82 2.4<br />

∗P


15 Mutualistic Interactions with Fungal Root Endophytes 269<br />

Methylobacterium spp. that synthesise cytokinin in vitro, are found endophytically<br />

in actively growing tissues of many or even all plants. They<br />

are present in the apoplast and could degrade metabolic wastes, producing<br />

e.g. ammonium ions, which could be metabolised by the host. Holland<br />

concluded that since endophyte-free plant tissue has not been conclusively<br />

shown to synthesise cytokinins, “microbial symbionts are not accidental<br />

visitors. They are co-evolved participants in plant physiology.” Similar signals<br />

seem to be generated by fungal endophytes, which not only synthesise<br />

the metabolites reported above (Kimura et al. 1992; Yates et al. 1997; Rey et<br />

al. 2001; Römmert et al. 2002), but also cytokinins (Petrini 1991; Tudzynski<br />

1997; Tudzynski and Sharon 2002).<br />

The plasticity of fungal root endophytes is demonstrated by the multiple<br />

options they have for regulating plant growth: synthesis of secondary<br />

metabolites and fungal phytohormones; directly providing nutrients, i.e.<br />

nitrogen and phosphate from the rhizosphere; but also bidirectional carbon<br />

transport. These results suggest that fungal endophytes are not only<br />

involved in a balanced antagonism with their hosts, but also may be responsible<br />

for a balanced regulation of growth. Thus, in at least some interactions,<br />

fungal root endophyte and host seem to be highly adapted to one<br />

another, resulting in a balanced interaction, as has also developed during<br />

the co-evolution of fungus and algae in lichens.<br />

15.5<br />

Disease Suppression<br />

Theideaofendophytemediatedinducedresistanceisanextensionof<br />

the defensive mutualism hypothesis, i.e. plants may acquire protection<br />

through constitutive and induced resistance (Bultman and Murphy 2000).<br />

As defined by Schönbeck et al. (1993), “induced resistance describes the<br />

improvement of natural resistance of a plant without alterations of the<br />

genome”. It is a non-specific general response that precludes absence of<br />

toxic effects for the parasite of the inducing agent as well as the absence of<br />

a dosage-response correlation, and it necessitates a time interval between<br />

application and response. Colonisation with fungal root endophytes has<br />

been reported to negatively influence growth of nematodes, insects and<br />

microbial pathogens (Selosse et al. 2004). However, it is not always clear<br />

whether induced resistance, altered plant nutrition, metabolism and/or<br />

sink effects are responsible for the increased mortality of predators when<br />

systemic effects are observed and mycotoxins are not involved.<br />

Systemic fungal root infections have been reported to lead to acquisition<br />

of induced resistance in both roots and/or shoots (Bargmann and Schönbeck<br />

1992; Hallmann and Sikora 1994; Sieber 2002). For example, root


270 B. Schulz<br />

inoculations of cabbage plants with soilborne endophytic Acremonium alternatum<br />

decreased damage due to larvae of the diamondback moth on the<br />

leaves (Raps and Vidal 1998). However, the authors suggested that competition<br />

for phytosterol, which both fungi and insects obtain from their host,<br />

rather than induced resistance could account for reduced larval growth on<br />

inoculated cabbage plants. They furthermore hypothesise that a disjunction<br />

of fungal colonisation and predation is responsible for the observed<br />

effect, the roots serving as a sink for the plant metabolite.<br />

Non-pathogenic Fusarium spp. have been found to induce resistance<br />

or afford the host cross-protection against pathogenic fungi (Kuldau and<br />

Yates 2000; Sieber 2002). For example, colonisation of the roots of pea and<br />

tomato by non-pathogenic strains of Fusarium oxysporum reduced disease<br />

by fungal root pathogens (Benhamou and Garand 2001) and predation<br />

by nematodes (Diedhiou et al. 2003), respectively. Fungal growth within<br />

the roots of pea was restricted to the outermost cell layers. Benhamou and<br />

Garand (2001) and Narisawa et al. (2004) found that resistance was induced,<br />

at least in part, by a massive elaboration of cell wall appositions and the<br />

deposition of electron-opaque material surrounding hyphae in the roots of<br />

pea and Chinese cabbage, respectively. This might be due to the accelerated<br />

deposition of lignin, a mechanical barrier for invading organisms, as was<br />

observed in shoots of maize when the roots were colonised by an avirulent<br />

isolate of F. moniliforme (Yates et al. 1997). Such host defence responses are<br />

initially produced to limit colonisation of the fungal endophytic invader,<br />

presumably resulting in a balance of antagonisms between host and fungus<br />

and in an asymptomatic endophytic infection. However, the host mechanical<br />

defence response also limits colonisation by subsequent pathogens.<br />

Only a few of the endophytic isolates from a given host are effective protectants<br />

when reinoculated into the host. Of 322 endophytic fungi isolated<br />

from Chinese cabbage (Brassica campestris), only 16 isolates, including Heteroconium<br />

chaetospira, Mortierella elongata, Westerdykella sp. and three<br />

non-sporulating isolates, almost completely suppressed disease caused by<br />

Plasmodiophora brassicae when reinoculated into the host in sterile soil<br />

(Narisawa et al. 1998; Usuki et al. 2002). The authors suggest that disease<br />

suppression by DSE may be strain specific, as also seems to be the case with<br />

growth enhancement (see Sect. 15.4). Other DSE have also been reported<br />

to protect roots against phytopathogenic fungi (Sieber 2002; Narisawa et<br />

al. 2002). For example, three endophytic fungi reduced disease caused by<br />

Verticillium longisporum in Chinese cabbage: two were Phialocephala and<br />

one an unidentified DSE (Narisawa et al. 2004). The latter DSE, which extensively<br />

colonised root cells of the cortex (in contrast to the P. fortinii<br />

isolates), was very effective in preventing disease even in field experiments.<br />

In addition, of 16 isolates of H. chaetospira that suppressed disease in axenically<br />

cultured seedlings, only 2 were also effective in non-sterile soil


15 Mutualistic Interactions with Fungal Root Endophytes 271<br />

(Narisawa et al. 2002). This could be due to the fact that in non-sterile soil<br />

the roots are naturally colonised by endophytes.<br />

Root colonisation with fungal endophytes can activate not only mechanical<br />

defence, but also induce the synthesis of defence metabolites in<br />

the roots, i.e. induce resistance. For example, colonisation of the roots of<br />

seedlings of Larix decidua with either the endophyte Cryptosporiopsis sp.<br />

or P. fortinii increased concentrations of soluble proanthocyanidins, and<br />

colonisation of barley roots with Fusarium sp. led to higher concentrations<br />

of phenylpropanoids in the roots than in those of the controls (Schulz et<br />

al. 1999). Similarly, Mucciarelli et al. (2003) found that the concentration of<br />

total phenolics increased significantly when Mentha piperita was colonised<br />

by a non-sporulating endophyte, though they found no significant change<br />

in the concentrations of total terpenoids.<br />

Picard et al. (2002) investigated what fungal factors were responsible<br />

for inducing systemic resistance when the mycoparasite Pythium oligandrum<br />

asymptomatically colonised the roots of tomato without inducing<br />

extensive cell damage. They demonstrated that colonisation led to cell wall<br />

appositions and to the deposition of phenolic metabolites (Benhamou et<br />

al. 1997). Picard et al. (2000) found that oligandrin, an elicitin-like protein<br />

produced by P. oligandrum, migrates within the vascular system of the host<br />

and induces systemic resistance.<br />

Improving plant resistance without the use of pesticides is one goal<br />

of biocontrol in agriculture. Only a small proportion of the fungal root<br />

endophytes of any one host seem capable of inducing systemic resistance.<br />

Are these endophytes present under natural environmental conditions at<br />

a sufficient colonisation density to induce resistance? In most cases, too<br />

little is presently known about what factor(s) induce resistance in the field.<br />

This is a broad and important field for future investigations.<br />

15.6<br />

Stress Tolerance<br />

Whereas mycorrhizal fungi can alleviate abiotic stresses, e.g. salt (Rabie<br />

2005) and osmotic stress (Ruiz-Lozano 2003), there have been only a few<br />

investigations studying the influence of endophytic colonisation on abiotic<br />

stress tolerance. One example strikingly demonstrates induction of<br />

stress tolerance to abiotic stress: a novel endophytic Curvularia sp., which<br />

colonised both roots and shoots of the host, increased host tolerance to<br />

temperatures of up to 65 ◦ C (Redman et al. 2002). Another abiotic stress<br />

is transplantation to polluted sites or micropropagation; stresses that can<br />

be counteracted by hardening. This was achieved not only by mycorrhization,<br />

but also by colonisation by the non-obligate biotrophic endophyte


272 B. Schulz<br />

Piriformospora indica (Sahay and Varma 1999). An arid climate also poses<br />

an abiotic stress, which DSE may help alleviate. Barrow and Aaltonen<br />

(2001) suggest that DSE may play such a role, because they found that<br />

under xeric, in contrast to mesic, conditions, colonisation by DSE was<br />

more prevalent than that by aseptate hyphae (AM). It is also possible that<br />

lipid accumulation within hyaline hyphae of DSE serves as an energyrich<br />

carbon reserve to sustain plants during extended drought (Barrow<br />

2003).<br />

The potential of fungal root endophytes to alleviate other abiotic and<br />

biotic stresses is a field that has not been adequately investigated. Their<br />

capacity to induce tolerance to other individual stressors, as well as to accumulated<br />

stress, should be tested both with individual and with a mixture<br />

of stressors at sublethal levels to determine the limits and potentials for<br />

induction of stress tolerance by fungal root endophytes, but also for future<br />

use in sustainable agriculture.<br />

15.7<br />

Factors Determining the Status of the Interaction<br />

Establishment of any interaction is always a multifactorial process. One<br />

factor that may contribute to the interaction being mutualistic is the pure<br />

extent of colonisation, since in the reported mutualistic interactions with<br />

fungal root endophytes, growth has been extensive (see Sect. 15.2.; Stone<br />

et al. 2000; Sieber 2002; Schulz and Boyle 2005). However, other factors e.g.<br />

biotic and abiotic stressors, nutrient support, and ontogenetic status, may<br />

also be relevant. As hypothesised by Schulz and Boyle (2005), mutualistic<br />

interactions have more frequently developed between microorganisms and<br />

the roots, because (1) the roots are a natural carbon sink of the plants and<br />

can supply dual and multi-organism symbioses with nutrients, (2) infection<br />

is less limited by xeromorphic structures, and (3) roots are in close contact<br />

with an environment harbouring many different, mainly degradatively<br />

active, micro-organisms.<br />

In spite of the examples of mutualistic interactions with fungal root<br />

endophytes presented in this chapter, it is important not to draw the conclusion<br />

that all interactions with fungal root endophytes are mutualistic.<br />

Only a small percentage of these endophytes seem to have the capability<br />

to interact mutualistically with their hosts (see Sects. 15.5, 15.6), perhaps<br />

because interactions with their hosts depend on the genetic predisposition<br />

and momentary metabolic status of the host, as well as on environmental<br />

factors. The same endophytes that under certain conditions interact mutualistically<br />

with their hosts may become pathogenic, for example when the<br />

host is stressed and the balance of the antagonism is tilted “in favour” of


15 Mutualistic Interactions with Fungal Root Endophytes 273<br />

the fungus (Kuldau and Yates 2000; Jumpponen 2001; Sieber 2002; Schulz<br />

and Boyle 2005).<br />

15.8<br />

Conclusions<br />

Fungal endophytic colonisation of the roots of plants is very variable (Table<br />

15.2), reflecting the plasticity of individual fungal endophytes, but<br />

also of endophytes as a whole. The hyphae can be hyaline and very thin,<br />

melanised, or develop microsclerotia. Colonisation is often extensive, may<br />

be inter- and/or intra-cellular, and is sometimes limited to the epidermis<br />

or cortex. The hyphae may extensively colonise the vascular cylinder and<br />

in particular the sieve elements, but also grow on the root surface and into<br />

the rhizosphere. Some DSE have even been observed to develop ericoid,<br />

ectendo- and ectomycorrhizal-like structures.<br />

Benefits for the fungal partner are a stable nutrient source and some<br />

protection from abiotic stresses. Factors that fungal root endophytes may<br />

contribute to a mutualistic interaction include secondary metabolites, phytohormones,<br />

nutrients, elicitins and colonisation (Fig. 15.2). These factors<br />

may have more than one benefit for the host. Potential advantages of the interactions<br />

for the host are summarised in Table 15.3 and include improved<br />

growth, induced resistance, stress tolerance, and protection from microbial<br />

and insect predators by mycotoxins.<br />

Morphologically and physiologically, endophytic root colonisations mirror<br />

the variability and thus the plasticity of endophytic interactions, but also<br />

Table 15.2. Endophytic plasticity<br />

Average shoot length (cm) Average dry weight of root+shoot (mg)<br />

Attribute In planta<br />

Colonisation Intercellular, intracellular, on/in: root surface, root hairs,<br />

epidermis, cortex, and/or vascular cylinder (xylem and<br />

phloem), systemic within roots, systemic within roots and<br />

shoots, local colonisation (?)<br />

Fungal morphologies Lipid-filled hyaline hyphae and protoplasts (?), melanised<br />

hyphae, microsclerotia, coiled hyphae, appressoria<br />

Fungal-plant associations No specialised structures Ericoid, ectendo-,<br />

and ectomycorrhizas<br />

Secondary metabolites Growth enhancement, growth inhibition, elicitation, balanced<br />

antagonism, microbial antagonism<br />

Phytohormones Modulate host growth<br />

Nutrient source Parasitically from host assimilates, saprophytically


274 B. Schulz<br />

Table 15.3. Endophytes/metabolites known to be involved in mutualistic interactions. DSE:<br />

Dark septate endophyte<br />

Advantage for host Endophyte / metabolite Host Reference<br />

Growth<br />

Cryptosporiopsis sp. Larix decidua Schulz et al. 2002<br />

enhancement Chaetomium, C. globosum Barley Vilich et al. 1998<br />

Cladorrhinum<br />

Cotton Gasoni and Stegman<br />

foecumdissimum<br />

de Gurfinkel 1997<br />

Fusarium Maize Yates et al. 1997<br />

Heteroconium chaetospira Chinese cabbage Usuki et al. 2002<br />

Oidiodendron maius Ericaceous plants Rice and Currah 2002<br />

Phialocephala fortinii Pinus contorta Jumpponen<br />

and Trappe 1998;<br />

Jumpponen et al.<br />

1998<br />

P. fortinii Larix decidua Schulz et al. 2002<br />

P. fortinii Carex firma, Haselwandter and<br />

C. curvula Read 1982<br />

Phomafimeti Vulpiaciliate Newsham 1994<br />

Stagonospora spp. Phragmites<br />

australis<br />

Ernst et al. 2003<br />

n.i. a Mentha piperita Mucciarelli et al.<br />

2002, 2003<br />

Altechromones A and B Lettuce Kimura et al. 1992<br />

Fumonisin Maize Yates et al. 1997<br />

Disease<br />

Acremonium Tomato Raps and Vidal 1998<br />

suppression/ Fusarium moniliforme Maize Yates et al. 1997<br />

induced resistance Fusarium oxysporum Pea Benhamou<br />

and Garand 2001<br />

F. oxysporum Tomato Diedhiou et al. 2003<br />

Fusarium spp. Various crops Kuldau and Yates<br />

2000<br />

H. chaetospira, Mortierella Chinese cabbage Narisawa et al. 1998;<br />

elongata, Westerdykella sp.<br />

Usuki et al. 2002<br />

P. fortinii Chinese cabbage Narisawa et al. 2004<br />

Pythium oligandrum Tomato Benhamou et al.<br />

1997; Picard et al.<br />

2002<br />

Oligandrin Tomato Picard et al. 2000<br />

Stress tolerance Curvularia sp. Dicanthelium<br />

lanuginosum<br />

Redman et al. 2002<br />

Piriformospora indica Tobacco Sahay and Varma<br />

1999<br />

DSE Atriplex Barrow and Aaltonen<br />

canescens 2001


15 Mutualistic Interactions with Fungal Root Endophytes 275<br />

Table 15.3. (continued)<br />

Advantage for host Endophyte / metabolite Host Reference<br />

Mycotoxins Cryptosporiopsis sp. Schulz et al. 1995<br />

Fusarium oxysporum Khambay et al. 2000<br />

Fusarium spp Chap. Bacon and<br />

Yates; Kuldau and<br />

Yates 2000;<br />

Miller 2001<br />

Verticillium<br />

Hallmann and Sikora<br />

chlamydosporium<br />

1996<br />

a Not identified<br />

Fig.15.2. Potential contributions of fungal root endophytes to mutualistic interactions,<br />

assuming extensive colonisation of the root. “Colonisation” indicates that it is unknown<br />

what aspect of fungal colonisation induces the response<br />

the evolutionary developmental stages from ubiquitous endophyte to mutualistic<br />

endophyte to specialised mycorrhizal fungus. Not only the exploitive,<br />

but also the mutualistic life history strategy becomes more prevalent with<br />

increasing specialisation (Brundrett 2002; see Chap. 16 by Brundrett).<br />

Acknowledgements. My thanks go to Anne-Kathrin Römmert and Miruna<br />

Oros-Sichler for permission to use unpublished results and to Christine<br />

Boyle for constructive discussions that helped develop the manuscript and<br />

ideas presented above.


276 B. Schulz<br />

<strong>References</strong><br />

Arnold AE, Herre EA (2003) Canopy cover and leaf age affect colonisation by tropical fungal<br />

endophytes: ecological pattern and process in Theobroma cacao (Malvaceae). Mycologia<br />

95:388–398<br />

Azevedo JL, Maccheroni W, Pereira JO, de Araujo WL (2000) Endophytic microorganisms:<br />

a review on insect control and recent advances on tropical plants. Electron J Biotechnol<br />

3:1–36<br />

Bacon CW, Hinton NS (1996) Symptomless endophytic colonisation of maize by Fusarium.<br />

Can J Bot 74:1195–1202<br />

Bargmann C, Schoenbeck F (1992) Acremonium kiliense as inducer of resistance to wilt<br />

disease on tomatoes. Z Pflanzenkr Pflanzenschutz 99:266–272<br />

Barrow JR (2003) Atypical morphology of dark septate fungal root endophytes of Bouteloua<br />

in arid southwestern USA rangelands. Mycorrhiza 13:239–247<br />

Barrow JR, Aaltonen RE (2001) Evaluation of the internal colonisation of Atriplex canescens<br />

(Pursh) Nutt. roots by dark septate fungi and the influence of host physiological activity.<br />

Mycorrhiza 11:199–205<br />

Benhamou N, Garand C (2001) Cytological analysis of defense-related mechanisms induced<br />

in pea root tissues in response to colonisation by nonpathogenic Fusarium oxysporum<br />

Fo47. Phytopathology 91:730–740<br />

Benhamou N, Rey P, Chérif M, Hockenhull J, Tirilly Y (1997) Treatment with the mycoparasite<br />

Pythium oligandrum triggers induction of defense-related reactions in tomato<br />

roots when challenged with Fusarium osysporum f. sp. radicis-lycopersici.Phytopathology<br />

87:108–122<br />

Boyle C, Götz M, Dammann-Tugend U, Schulz B (2001) Endophyte - host interactions III.<br />

Local vs. systemic colonisation. Symbiosis 31:259–281<br />

Brundrett MC (2002) Tansley Review no. 134: Coevolution of roots and mycorrhizas of land<br />

plants. New Phytol 154:275–304<br />

Bultman TL, Murphy JC (2000) Do fungal endophytes mediate wound-induced resistance?<br />

In: Bacon CW, White JF (eds) Microbial endophytes. Dekker, New York, pp 421–455<br />

Caldwell BA, Jumpponen A, Trappe JM (2000) Utilization of major detrital substrates by<br />

dark-septate, root endophytes. Mycologia 92:230–232<br />

Danielson RM, Visser S (1990) The mycorrhizal and nodulation status of container-grown<br />

trees and shrubs reared in commercial nurseries. Can J For Res 20:609–614<br />

Demain AL (1980) Do antibiotics function in nature? Search 11:148–151<br />

Diedhiou PM, Hallmann J, Oerke EC (2003) Effects of arbuscular mycorrhizal fungi and<br />

a non-pathogenic Fusarium oxysporum on Meloidogyne incognita infestation of tomato.<br />

Mycorrhiza 13:199–204<br />

Ernst M, Mendgen KW, Wirsel SG (2003) Endophytic fungal mutualists: seed-borne<br />

Stagonospora spp. enhance reed biomass production in axenic microcosms. Mol Plant-<br />

Microbe Interact 16:580–587<br />

Fang-ting W, Dai-jie C, Xiu-ping Q (2004) Recent studies of bioactive substances produced<br />

by endophyte. Chin J Antibiot 29:184–192<br />

Fernando AA, Currah RS (1996) A comparative study of the effects of the root endophytes<br />

Leptodontidium orchidicola and Phialocephala fortinii (Fungi imperfecti) on the growth<br />

of some subalpine plants in culture. Can J Bot 74:1071–1078<br />

Gasoni L, Stegman De Gurfinkel B (1997) The endophyte Cladorrhinum foecundissimum in<br />

cotton roots: phosphorus uptake and host growth. Mycol Res 101:867–870<br />

Hallmann J, Sikora RA (1994) Influence of Fusarium oxysporum, a mutualistic fungal endophyte<br />

on Meloidogyne javanica infection of tomato. J Plant Dis Prot 101:475–481


15 Mutualistic Interactions with Fungal Root Endophytes 277<br />

Hallmann J, Sikora RA (1996) Toxicity of fungal endophyte secondary metabolites to<br />

plant-parasitic nematodes and soil-borne plant-pathogenic fungi. Eur J Plant Pathol<br />

102:155–162<br />

Haselwandter K, Read DJ (1982) The significance of a root-fungus association in two Carex<br />

species of high-alpine plant communities. Oecologia 53:352–254<br />

Hashimoto Y, Hyakumachi M (2001) Effects of isolates of ectomycorrhizal fungi and endophytic<br />

Mycelium radicis atrovirens that were dominant in soil from disturbed sites on<br />

growth of Betula platyphylla var. japonica seedlings. Ecol Res 16:117–125<br />

Holland MA (1997) Occam’s razor applied to hormonology. Are cytokinins produced by<br />

plants? Plant Physiol 115:865–868<br />

Jallow JFA, Digassa-Gobena D, Vidal S (2004) Indirect interaction between an unspecialized<br />

endophytic fungus and a polyphagous moth. Basic Appl Ecol 5:183–191<br />

Jumpponen A (1999) Spatial distribution of discrete RAPD phenotypes of a root endophytic<br />

fungus, Phialocephala fortinii, at a primary successional site on a glacier forefront. New<br />

Phytol 141:333–344<br />

Jumpponen A (2001) Dark septate endophytes – are they mycorrhizal? Mycorrhiza<br />

11:207–211<br />

Jumpponen A, Trappe JM (1998) Performance of Pinus contorta inoculated with two strains<br />

of root endophytic fungus, Phialocephala fortinii: effects of synthesis system and glucose<br />

concentration. Can J Bot 76:1205–1213<br />

Jumpponen A, Mattson KG, Trappe JM (1998) Mycorrhizal functioning of Phialocephala<br />

fortinii with Pinus contorta on glacier forefront soil: interactions with soil nitrogen and<br />

organic matter. Mycorrhiza 7:261–265<br />

Kaldorf M, Renker C, Fladung M, Buscot F (2004) Characterization and spatial distribution of<br />

ectomycorrhizas colonising aspen clones released in an experimental field. Mycorrhiza<br />

14:295–306<br />

Khambay BPS, Bourne JM, Carmeron S, Kerry BR, Zaki MJ (2000) A nematicidal metabolite<br />

from Verticillium chlamydosporium. Pest Manage Sci 56:1098–1099<br />

Kimura Y, Mizuno T, Nakajima H, Hamasaki T (1992) Altechromones A and B, new plant<br />

growth regulators produced by the fungus Alternaria sp. Biosci Biotechnol Biochem<br />

56:1664–1665<br />

Kuldau GA, Yates IE (2000) Evidence for Fusarium endophytes in cultivated and wild plants.<br />

In: Bacon CW, White JF (eds) Microbial endophytes. Dekker, New York, pp 85–120<br />

Lubuglio KF, Wilcox HE (1988) Growth and survival of ectomycorrhizal and ectoendomycorrhizal<br />

seedlings of Pinus resinosa on iron tailings. Can J Bot 66:55–60<br />

Miller JD (2001) Factors that affect the occurrence of fumonisin. Environ Health Perspect<br />

109:321–324<br />

Miller JD, Mackenzie S, Foto M, Adams GW, Findlay JA (2002) Needles of white spruce<br />

inoculated with rugulosin-producing endophytes contain rugulosin reducing spruce<br />

budworm growth rate. Mycol Res 106:471–479<br />

Mucciarelli M, Scannerini S, Bertea C, Maffei M (2002) An ascomycetous endophyte isolated<br />

from Mentha piperita L.: biological features and molecular studies. Mycologia<br />

94:28–39<br />

Mucciarelli M, Scannerini S, Bertea C, Maffei M (2003) In vitro and in vivo peppermint<br />

(Mentha piperita) growth promotion by nonmycorrhizal fungal colonisation. New Phytol<br />

158:579–591<br />

Narisawa K, Tokumasu S, Hashiba T (1998) Suppression of clubroot formation in Chinese<br />

cabbage by the root endophytic fungus, Heteroconium chaetospira. Plant Pathol<br />

47:206–210<br />

Narisawa K, Kawamata H, Currah RS, Hashiba T (2002) Suppression of Verticillium wilt in<br />

eggplant by some fungal root endophytes. Eur J Plant Pathol 108:103–109


278 B. Schulz<br />

Narisawa K, Usuki F, Hashiba T (2004) Control of Verticillium yellows in Chinese cabbage<br />

by the dark septate endophytic fungus LtVB3. Phytopathology 94:412–418<br />

Newsham KK (1994) First record of intracellular sporulation by a coelomycete fungus. Mycol<br />

Res 98:1390–1392<br />

Ohki T, Masuya H, Yonezawa M, Usuki F, Narisawa, K, Hashiba T (2002) Colonisation process<br />

of the root endophytic fungus Heteroconium chaetospira in roots of Chinese cabbage.<br />

Mycoscience 43:191–194<br />

Petrini O (1991) Fungal endophytes of tree leaves. In: Andrews J, Hirano S (eds) Microbial<br />

ecology of leaves. Springer, New York Berlin Heidelberg, pp 179–197<br />

Picard K, Ponchet M, Blein J-P, Rey P, Tirilly Y, Benhamou N (2000) Oligandrin. A proteinaceous<br />

molecule produced by the mycoparasite Pythium oligandrum induces resistance<br />

to Phytophthora parasitica infection in tomato plants. Plant Physiol 124:379–395<br />

Rabie GH (2005) Influence of arbuscular mycorrhizal fungi and kinetin on the response of<br />

mungbean plants to irrigation with seawater. Mycorrhiza 15:225–230<br />

Raps A, Vidal S (1998) Indirect effects of an unspecialized endophytic fungus on specialized<br />

plant-herbivorous insect interactions. Oecologia 114:541–547<br />

Redman RS, Sheehan KB, Stout RG, Rodriguez RJ, Henson JM (2002) Thermotolerance<br />

generated by plant/fungal symbiosis. Science 298:1581<br />

Rey P, Leucart S, Désilets H, Bélanger RR, Larue JP, Tirilly Y (2001) Production of indole-<br />

3-acetic acid and tryptophol by Pythium group F: possible role in pathogenesis. Eur<br />

J Plant Pathol 107:895–904<br />

Rice AV, Currah RS (2002) New perspectives on the niche and holomorph of the myxotrichoid<br />

hyphomycete, Oidodendron maius. Mycol Res 106:1463–1467<br />

Römmert AK, Oros-Sichler M, Aust H-J, Lange T, Schulz B (2002) Growth promoting effects<br />

of endophytic colonisation of the roots of larch (Larix decidua) withCryptosporiopsis<br />

sp. and Phialophora sp. 7th International Mycological Congress, Oslo, Norway, p 309<br />

Ruiz-Lozano JM (2003) Arbuscular mycorrhizal symbiosis and alleviation of osmotic stress.<br />

New perspectives for molecular studies. Mycorrhiza 13:309–317<br />

Sahay NS, Varma A (1999) Piriformospora indica: a new biological hardening tool for<br />

micropropagated plants. FEMS Microbiol Lett 181:297–302<br />

Schönbeck F, Steiner U, Kraska T (1993) Induzierte Resistenz: Kriterien, Mechanismen,<br />

Anwendung und Bewertung. J Plant Dis Prot 100:541–557<br />

Schulz B, Boyle C (2005) The endophytic continuum. Mycol Res 109:661–687<br />

Schulz B, Sucker J, Aust H-J, Krohn K, Ludewig K, Jones PG, Döring D (1995) Biologically<br />

active secondary metabolites of endophytic Pezicula species. Mycol Res 99:1007–1015<br />

Schulz B, Römmert A-K, Dammann U, Aust H-J, Strack D (1999) The endophyte-host<br />

interaction: a balanced antagonism. Mycol Res 103:1275–1283<br />

Schulz B, Boyle C, Draeger S, Römmert A-K, Krohn K (2002) Endophytic fungi: a source of<br />

biologically active secondary metabolites. Mycol Res 106:996–1004<br />

Selosse MA, Baudoin E, Vandenkoorhuyse P (2004) Symbiotic microorganisms, a key for<br />

ecological success and protection of plants. C R Biol 327:639–648<br />

Sen R, Hietala AM, Zelmer CD (1999) Common anastomosis and internal transcribed<br />

spacer RFLP groupings in binucleate Rhizoctonia isolates representing root endophytes<br />

of Pinus sylvestris, Ceratorhiza spp. from orchid mycorrhizas and a phytopathogenic<br />

anastomosis group. New Phytol 44:331–341<br />

Sieber TN (2002) Fungal root endophytes In: Waisel Y, Eshel A, Kafkafi U (eds) The hidden<br />

half. Dekker, New York, pp 887–917<br />

Sieber TN, Riesen TK, Müller E, Fried PM (1988) Endophytic fungi in four winter wheat<br />

cultivars (Triticum aestivum L.) differing in resistance against Stagonospora nodorum<br />

(Berk.) Cast. & Germ. = Septorianodorum (Berk.) Berk. J Phytopathol 122:289–306


15 Mutualistic Interactions with Fungal Root Endophytes 279<br />

Stone JK, Bacon CW, White JF (2000) An overview of endophytic microbes: endophytism<br />

defined. In: Bacon CW, White JF (eds) Microbial endophytes. Dekker, New York, pp 3–30<br />

Tan RX, Zou WX (2001) Endophytes: a rich source of functional metabolites. Nat Prod Rep<br />

18:448–459<br />

Tudzynski B (1997) Fungal phytohormones in pathogenic and mutualistic associations. In:<br />

Carroll GC, Tudzynski P (eds) The mycota V. Springer, Berlin Heidelberg New York,<br />

pp 167–184<br />

Tudzynski B, Sharon A (2002) Biosynthesis, biological role and application of fungal phytohormones<br />

In: Osiewacz HD (ed) The mycota X. Industrial applications. Springer, Berlin<br />

Heidelberg New York, pp 183–212<br />

Usuki F, Narisawa K (2005) Formation of structures resembling ericoid mycorrhizas by<br />

the root endophytic fungus Heteroconium chaetospira within roots of Rhododendron<br />

obtusum var. kaempferi. Mycorrhiza 15:61–64<br />

Usuki F, Narisawa K, Yonezawa M, Kakishima M, Hashiba T (2002) An efficient method<br />

for colonisation of Chinese cabbage by the root endophytic fungus Heteroconium<br />

chaetospira. J Gen Plant Pathol 68:326–332<br />

Varma A, Singh A, Sahay NS, Sharma J, Roy A, Kumari M, Raha D, Thakran S, Deka D,<br />

Bharti K, Hurek T, Blechert O, Rexer K-H, Kost G, Hahn A, Maier W, Walter M, Strack D,<br />

Kranner I (2000) Piriformospora indica: an axenically culturable mycorrhiza-like endosymbiotic<br />

fungus In: Hock B (ed) The mycota, vol IX Fungal associations. Springer,<br />

Berlin Heidelberg New York, pp 125–150<br />

Vilich V, Dolfen M, Sikora RA (1998) Chaetomium spp. colonisation of barley following<br />

seed treatment and its effect on plant growth and Erysiphe graminis f. sp. hordei disease<br />

severity. Z Pflanzenkr Pflanzenschutz 105:130–139<br />

Vrålstad T, Schumacher T, Taylor AFS (2002) Mycorrhizal synthesis between fungal strains of<br />

the Hymenoscyphus ericae aggregate and potential ectomycorrhizal and ericoid hosts.<br />

New Phytol 153:143–152<br />

Yates IE, Bacon CW, Hinton DM (1997) Effects of endophytic infection by Fusarium moniliforme<br />

on corn growth and cellular morphology. Plant Dis 81:723–728


16<br />

Understanding the Roles<br />

of Multifunctional Mycorrhizal<br />

and Endophytic Fungi<br />

Mark C. Brundrett<br />

16.1<br />

How Mycorrhizal Fungi Differ from Endophytes<br />

16.1.1<br />

Definitions<br />

The most appropriate definition of endophytism is that of symptomless<br />

associations by organisms that grow within living plant tissues (see other<br />

chapters in this book). Many fungi colonise the cortex of living roots without<br />

causing disease, including pathogenic or necrotrophic fungi with latent<br />

phases as well as beneficial fungi that offer protection against pathogens,<br />

but it is difficult to precisely categorise these fungi [Saikkonen et al. 1998;<br />

Sivasithamparam 1998; see Chaps. 1 (Schulz and Boyle) and 8 (Bacon and<br />

Yates)]. A new definition of mycorrhizal associations was required to adequately<br />

separate them from other root-fungus associations (Brundrett<br />

2004). According to this definition, endophytic associations differ from<br />

mycorrhizas primarily by the absence of a localised interface of specialised<br />

hyphae, the absence of synchronised plant-fungus development, and the<br />

lack of plant benefits from nutrient transfer – the three key defining features<br />

of mycorrhizas. However, plants may benefit indirectly from endophytes<br />

by increased resistance to herbivores, pathogens or stress, or by other unknown<br />

mechanisms, as described elsewhere in this book (see in particular<br />

Chap. 15 by Schulz).<br />

There are a number of distinct categories of mycorrhizal fungi, all of<br />

which differ from other fungi primarily because they are dual soil-plant<br />

inhabitants that are efficient at growth and nutrient uptake in both soils and<br />

plants (Brundrett 2002). Conversely, endophytes and pathogens are plant<br />

inhabitants, which do not require efficient means of acquiring nutrients<br />

from soils to supply plants (Table 16.1). The examples in this chapter<br />

primarily concern root-fungus associations, but mycorrhizal fungi and<br />

Mark C. Brundrett: School of Plant Biology, Faculty of Natural and Agricultural Sciences,<br />

The University of Western Australia, Crawley, WA 6009, Australia, E-mail:<br />

mark.brundrett@environmnent.wa.gov.au<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


282 M.C. Brundrett<br />

Table 16.1. Comparison of mycorrhizal, parasitic and endophytic root associations (Brundrett<br />

2004). VAM Vesicular arbuscular mycorrhizas, ECM ectomycorrhizas<br />

Criteria Mycorrhizal Parasitic Endophytic<br />

Morphology Specialised hyphae in<br />

specialised plant organ<br />

Specialised hyphae Relatively<br />

unspecialised hyphae<br />

Development Synchronised Often synchronised Not synchronised<br />

Impact on fungus Obligate requirement<br />

for plant supplied<br />

nutrients (VAM and<br />

ECM)<br />

Fungus obligately or<br />

facultatively<br />

dependent on plant<br />

Fungus moderately or<br />

weakly dependent<br />

Impact on plant Strong or weak benefit Strong or weak harm Weak harm or benefit<br />

Nutrient transfer Synchronised transfer,<br />

fungusastrongsink<br />

Active or passive<br />

transfer, fungus<br />

astrongorweaksink<br />

Passive transfer,<br />

fungus not a strong<br />

sink<br />

endophytes also occupy other substrate-contacting organs (e.g. rhizomes,<br />

stems, scale leaves, etc.).<br />

16.1.2<br />

Roles of Endophytes and Mycorrhizal Fungi<br />

It is harder to separate mycorrhizal fungi from other functional groups<br />

of fungi than it is to separate mycorrhizal associations from other plantfungus<br />

relationships, as mycorrhizas are defined by morphological criteria<br />

(Brundrett 2004). Differences between endophytic associations and mycorrhizal<br />

associations are summarised in Table 16.1. The most important<br />

of these differences are the absence of substantial fungus-to-plant nutrient<br />

transfer in endophytic associations and the lack of synchronised development<br />

between endophytes and plants. Endophytic associations lack<br />

the specialised interface, formed by complex hyphal growth synchronised<br />

with substantial cytoplasm synthesis by host cells, that is diagnostic of<br />

mycorrhizas. The complex host-fungus interface in mycorrhizal associations<br />

allows rapid bi-directional transfer of nutrients over a relatively short<br />

time (a few weeks). However, nutrient transfer to fungi in endophytes is<br />

still likely to be important to the fungus if it continues over much longer<br />

time-spans (months or years) than the active phase of mycorrhizal associations.<br />

We would also not expect plant-fungus coevolution in endophytic<br />

associations if plants do not receive substantial direct benefits from these<br />

organisms. However, the results of plant evolution to become more efficient<br />

at forming mycorrhizas (e.g. roots that evolved to house fungi, Brundrett<br />

2002) would ultimately have also benefitted endophytes.


16 Roles of Endophytic and Mycorrhizal Fungi 283<br />

While we can safely say that endophytes are not mycorrhizal (as defined<br />

above), it seems likely that mycorrhizal fungi have endophytic phases<br />

involving long-term coexistence with plants without active growth or nutrient<br />

transfer. For example, old roots are an important source of inoculum<br />

for Glomeromycete fungi, which can survive as endophytes in living roots<br />

for up to 10 years after arbuscules collapse (Brundrett and Kendrick 1988).<br />

These fungi can also have a necrotrophic phase in dying roots, providing<br />

thefunguswithfirstaccesstonutrientsthatcanbetransferredtohyphae<br />

within other plants (Eason et al. 1991). Examples of endophytic activity by<br />

mycorrhizal fungi are summarised in the next section.<br />

16.2<br />

Endophytic Activity by Mycorrhizal Fungi<br />

Endophytic growth of mycorrhizal fungi in plants is common and differs<br />

from mycorrhizal associations primarily by the absence of defining<br />

morphological features, as discussed above. It is not difficult to explain<br />

the endophytic competence of mycorrhizal fungi in non-host plants, as<br />

we would expect that they would normally have the highest inoculum potentials<br />

of soil organisms and these fungi require the capacity to rapidly<br />

and efficiently colonise plant roots. It has been suggested that plants must<br />

employ effective defences if they are to remain free of mycorrhizal fungi<br />

(Brundrett 1991).<br />

Another difference between mycorrhizal and endophytic fungi is that the<br />

nutritional resources provided by endophytic activity (access to endophytic<br />

exudates) do not seem to be sufficient to sustain mycorrhizal fungi, which<br />

generally cannot persist in soils without forming mycorrhizas with host<br />

plants (Brundrett 1991). It has been suggested that mycorrhizal fungi may<br />

endophytically colonise roots and other soil organisms for shelter rather<br />

than food, to avoid the many soil organisms that prey on them (Koske<br />

1984). Such use of roots as shelters by mycorrhizal fungi likely would be<br />

of limited immediate significance to the plant in which this occurs, but<br />

should ultimately benefit other plants in the same ecosystem, by maintaining<br />

a reservoir of inoculum. Examples of suggested endophytic associations<br />

by mycorrhizal fungi and other fungi that frequent mycorrhizas are listed<br />

in sections below, with fungi classified by their primary roles.<br />

16.2.1<br />

Glomeromycotan (Vesicular-Arbuscular Mycorrhizal) Fungi<br />

Glomeromycotan fungi, forming vesicular arbuscular mycorrhizas [VAM,<br />

or arbuscular mycorrhizas (AM)], are ubiquitous soil organisms that


284 M.C. Brundrett<br />

proliferate within patches of soil organic material (St John et al. 1983; Joner<br />

and Jakobsen 1995; Azcón-Aguilar et al. 1999). As shown in Fig. 16.1A, B,<br />

these fungi commonly grow in non-host plants (Brundrett and Kendrick<br />

1988; Imhof 2001). They also occupy plant organs other than roots, dead<br />

soil animals and spores of other VAM fungi, presumably to acquire nutri-<br />

Fig.16.1. A–D Roots cleared in potassium hydroxide and stained with Chlorazol Black E in<br />

lactoglycerol. A, B Endophytic growth of a Glomeromycotan fungus in the nonmycorrhizal<br />

plant Hydrophyllum virginianum consisting of vesicles (v)andhyphae(arrow)inarhizome<br />

scale A and nonmycorrhizal root B. C, D Dense growth by unknown fungi (arrows)onthe<br />

epidermis of roots of Scaevola crassifolia C and Acer saccharum D. These plants also have<br />

vesicular-arbuscular mycorrhizas. E, F Unstained roots of Eucalyptus globulus grown in<br />

pasteurised soil colonised by an unidentified opportunistic conidial fungus. E Hyphae and<br />

conidia (arrows) formed in soil. F Hyphae on the surface of long roots forming a patchy<br />

Hartig-net-like structure (arrow)


16 Roles of Endophytic and Mycorrhizal Fungi 285<br />

ents or avoid predation (Rabatin and Rhodes 1982; Warner 1984; St John<br />

et al. 1983; Koske 1984; Brundrett and Kendrick 1988). Mycorrhizal colonisation<br />

of some mutants of normally mycorrhizal species also resembles<br />

endophytic activity by these fungi, as they form hyphae and vesicles but<br />

not arbuscles (Demchenko et al. 2004). This suggests that VAM fungi will<br />

switch between endophytic and mycorrhizal activity in plants in response<br />

to signals provided by the plant.<br />

Nonmycorrhizal plants are defined as those that normally exclude mycorrhizal<br />

fungi from their healthy young roots (Brundrett 1991). As implied<br />

by this definition, endophytic colonisation of nonmycorrhizal plant roots<br />

byVAMfungiiscommoninolderroots,butisconsideredtobeoflimited<br />

functional significance because it does not result in plant growth responses<br />

(Ocampo 1986; Muthukumar et al. 1997; Giovannetti and Sbrana 1998).<br />

However, there may be a fine line between nonmycorrhizal and facultatively<br />

mycorrhizal species in which VAM associations are present or absent<br />

as a result of soil conditions (see Brundrett 1991). These issues are best illustrated<br />

by the debate about the role of mycorrhizal associations in sedges<br />

(Cyperaceae and allied families) that has been ongoing for decades (Powell<br />

1975; Tester et al. 1987; Brundrett 1991; Muthukumar et al. 2004). The<br />

statement by Muthukumar et al. (2004) that the Cyperaceae is a mycorrhizal<br />

family, is in disagreement with the opinion of earlier reviewers (e.g.<br />

Powell 1975; Tester et al. 1987; Brundrett 1991) and is not as clear-cut as it<br />

seems. Half of sedges examined in studies summarised by Muthukumar et<br />

al. (2004) contained mycorrhizal fungi. However, they could not clearly distinguish<br />

endophytic from mycorrhizal associations using the information<br />

provided in many studies, and they suspected that these associations were<br />

often non-functional. Detailed studies of sedges such as Carex spp. have<br />

found them to be nonmycorrhizal throughout the year with endophytic<br />

VAM hyphae in older roots (Brundrett and Kendrick 1988; Cornwell et<br />

al. 2001). There may be a continuum from mycorrhizal to nonmycorrhizal<br />

species in plant families such as the Cyperaceae. In facultatively mycorrhizal<br />

species, VAM fungi may sometimes function more as endophytes than as<br />

mycorrhizal associates, as they can contribute to disease suppression in the<br />

absence of growth responses (Newsham et al. 1995; Cordier et al. 1998).<br />

It is often assumed that sparse or inconsistent mycorrhizal formation is<br />

of limited benefit to plants, but this has rarely been tested in experiments<br />

using appropriate soil conditions.<br />

A second common example of cases where the hyphal growth of Glomeromycotan<br />

fungi is difficult to interpret is their colonisation of roots of<br />

ectomycorrhizal (ECM) plants in species with dual mycorrhizal associations.<br />

In plants with both ECM and VAM, their relative importance can<br />

vary due to the age of plants and the habitats in which they grow (Chen<br />

et al. 2000; van der Heijden 2001). Most trees with dual ECM/VAM asso-


286 M.C. Brundrett<br />

ciations are considered to benefit primarily from ECM, but large growth<br />

responses to VAM occur in experiments, especially when plants are young<br />

(Brundrett et al. 1996; Chen et al. 2000). A further complication is the fact<br />

that Glomeromycete fungi are commonly present as endophytes in roots of<br />

ECM plants (e.g. Harley and Harley 1987; Cázares and Trappe 1993; Smith,<br />

et al. 1998). This endophytic activity is distinguished from functional dual<br />

ECM/VAM associations by the absence of arbuscules. Glomeromycotan<br />

fungi can also grow endophytically in orchids (Hall 1976).<br />

16.2.2<br />

Ectomycorrhizal Fungi<br />

Ectomycorrhizal associations are defined by a key morphological criterion:<br />

labyrinthine Hartig net hyphae in an interface between cells of the primary<br />

root cortex or epidermis. However, designation of ECM without a welldefined<br />

Hartig net is not always clear cut, as “superficial” ECM roots with<br />

a thin or patchy Hartig net occur in synthesis experiments using hostfungus<br />

combinations that are not fully compatible (Burgess et al. 1994;<br />

Peterson and Massicotte 2004). These also occur in nature, where they are<br />

believed to benefit plants (Malajczuk et al. 1987). When initiated by spores<br />

or other limited sources of inoculum in experiments, ECM fungi weakly<br />

colonise root surfaces before they have sufficient energy to form typical<br />

mycorrhizas (Chilvers and Gust 1982). This establishment phase may equate<br />

to a transition from endophyte-like to mutualistic activity. Some fungi form<br />

associations with a mantle but no Hartig net on non-host roots, such as<br />

those of Morchella sp. on Pinaceae (Dahlstrom et al. 2000), Cortinarius sp.<br />

on Carex (Harrington and Mitchell 2002) and Tricholoma sp. on Pinus (Gill<br />

et al. 1999). The opportunistic growth of fungal hyphae on roots is common<br />

in nature and could be considered a form of endophytism in which fungi<br />

feed on root exudates without penetrating cells (Fig. 16.1C, D).<br />

Other examples of fungi that seem to occupy intermediate positions<br />

on an ECM-endophyte continuum are opportunistic fungi that weakly<br />

colonise seedling roots in the nursery (see Fig. 16.1E, F). These nursery<br />

fungi fill an empty niche in potting mixes that lack mycorrhizal fungus<br />

propagules or that are not conducive to growth by mycorrhizal fungi.<br />

These nursery fungi are probably of limited functional significance, as<br />

suggested by similar fungal growth on both long and short roots, a patchy<br />

Hartig net and the absence of morphological responses by short roots<br />

(root swelling). Ectendomycorrhizas, a variant of ECM, also tend to occur<br />

in substrates lacking other fungi and where nutrient levels are high, and so<br />

are considered to be of limited benefit to plants (Yu et al. 2001).


16 Roles of Endophytic and Mycorrhizal Fungi 287<br />

16.2.3<br />

Fungi in Orchids<br />

Most fungi that form orchid mycorrhizas are basidiomycetes belonging to<br />

a diverse assemblage of fungi assigned by morphological features to the<br />

asexual form genus Rhizoctonia (see also Chap. 9 by Bayman and Otero).<br />

These fungi are also more precisely defined using anastomosis reactions,<br />

enzyme assays and DNA-based methods (Currah et al. 1997; Taylor et al.<br />

2002; McKormick et al. 2004). Mycorrhizal associates of orchids occur in all<br />

three clades of the polyphyletic Rhizoctonia complex (Sebacinaceae, Tulasnellales,<br />

Ceratobasidiales), but are rare in sister clades of basidiomycetes<br />

(Taylor et al. 2002). As shown in Table 16.2, Sebacina members include<br />

ECM fungi, orchid mycorrhizal associates, and bryophyte fungi, and also<br />

occur in ericoid plants, but there is some doubt about how much roles<br />

overlap between separate clades within this genus (Warcup 1981; Selosse et<br />

al. 2002a, 2002b). Another multifunctional orchid fungus is Thanatephorus<br />

gardneri, the mycorrhizal associate of the underground orchid (Rhizanthella<br />

gardneri), which also forms ECM with Melaleuca uncinata and is<br />

capable of endophytic activity (Table 16.2). This multifunctional fungus is<br />

the only known member of the Rhizoctonia alliance outside of Sebacina<br />

that forms ECM.<br />

Attempts to isolate mycorrhizal fungi from orchids often produce cultures<br />

of bacteria, actinomycetes and common endophytes such as Fusarium<br />

and Verticillium,aswellasECMfungiandericoidfungi(Richardsonand<br />

Currah 1995; Currah et al. 1997; Kristiansen et al. 2001; Bayman et al. 2002;<br />

Otero et al. 2002, Bidartondo et al. 2004). Most of these other fungi seem<br />

to be endophytes, but Fusarium isolates are also capable of forming orchid<br />

mycorrhizas (Vujanovic et al. 2000; Abdul Karim 2005). We have observed<br />

that the frequency of isolation of non-Rhizoctonia fungi decreases substantially<br />

when fungi are isolated from single pelotons rather than from<br />

surface-sterilised tissue blocks, providing further evidence that many fungi<br />

isolated by the latter method are endophytes. It is recommended that only<br />

isolates that germinate orchids to an advanced seedling stage with a green<br />

leaf be designated as orchid mycorrhizal fungi (Batty et al. 2002).<br />

Kottke et al. (2003) found members of the Rhizoctonia complex (Tulasnella<br />

and Sebacina) closely related to orchid fungi that formed mycorrhizalike<br />

associations with hepatics. These bryophytes also contained ascomycetes,<br />

and the nature of these associations requires further investigation.<br />

Members of the Rhizoctonia complex include major pathogens of seedlings<br />

of crop and horticultural species (Sivasithamparam 1998), but the degree<br />

of overlap between pathogens and orchid associates is unknown (Batty et<br />

al. 2002). In general, fungi associating with orchids differ substantially<br />

from other mycorrhizal fungi, because they do not belong to discrete


288 M.C. Brundrett<br />

Table 16.2. Examples of fungi with multifunctional roles. Fungi within a row have been shown to be closely related by molecular identification<br />

Orchid mycorrhizas Other role<br />

Fungus Endophyte Pathogen Ectomycorrhizal Ericoid mycorrhizas<br />

Decomposition of<br />

wood?<br />

Myco-heterotrophic<br />

orchids<br />

(McCormick et al. 2004;<br />

Primary role<br />

(Köljalg et al. 2000)<br />

Tomentella spp.<br />

(Thelephoraceae)<br />

Taylor et al. 2002)<br />

Mycorrhizas in hepatics<br />

(Kottke et al. 2003) a .<br />

See below Many orchid associates<br />

(see text)<br />

Major role:<br />

(see text)<br />

Rhizoctonia complex S.L. Common: e.g.<br />

Pinus sylvestris<br />

Saprophytic<br />

(Sen et al. 1999)<br />

Myco-heterotrophic<br />

orchids (Selosse et al.<br />

2002b; Taylor et al.<br />

2003) a .Saprophytic?<br />

Some orchids (Warcup<br />

1981; Y. Bonnardeaux,<br />

personal<br />

Suspected<br />

role? (Allen<br />

et al. 2003)<br />

(Glen et al. 2002;<br />

Selosse et al. 2002a;<br />

Urban et al. 2003) a<br />

Ericaceae?<br />

(Allen et al. 2003)<br />

Sebacina spp.<br />

(in Rhizoctonia complex)<br />

communication)<br />

Saprophytic?<br />

Myco-heterotrophic<br />

orchid Rhizanthella<br />

gardneri (Warcup 1985;<br />

Mursidawati 2003)<br />

Melaleuca uncinata<br />

and other plants<br />

(Warcup 1985;<br />

Mursidawati 2003)<br />

ECM of trees<br />

(Harney et al. 1997)<br />

Some endophytic<br />

activity<br />

(Mursidawati<br />

Thanatephorus gardneri<br />

(in Rhizoctonia complex)<br />

Saprophytic phases<br />

Tropical orchids<br />

(Bayman et al. 2002;<br />

Abdul Karim 2005)<br />

e.g.<br />

Pamphile<br />

and Azevedo<br />

2002<br />

2003)<br />

Phialocephala spp. Widespread<br />

(DSE fungi)<br />

(see text)<br />

Fusarium spp. Common (Kuldau<br />

and Yates 2000;<br />

Redman et al.<br />

2001)


16 Roles of Endophytic and Mycorrhizal Fungi 289<br />

Table 16.2. (continued)<br />

Other role<br />

Orchid<br />

mycorrhizas<br />

Mycorrhizas of<br />

bryophytes (Duckett<br />

and Read 1995) a<br />

Fungus Endophyte Pathogen Ectomycorrhizal Ericoid<br />

mycorrhizas<br />

Hymenoscyphus ericae Picea mariana roots<br />

Six tree spp.<br />

(Piercey et al. 2002),<br />

(Vrålstad et al. 2000)<br />

Bryophyte<br />

a<br />

Primary role<br />

(see text)<br />

(Chambers et al. 1999)<br />

Some isolates are<br />

strong saprophytes<br />

(Piercey et al. 2002)<br />

Primary role?<br />

(see text)<br />

Oidiodendron spp. ECM roots of Quercus ilex<br />

(Bergero et al. 2000)<br />

Trees<br />

(Sakakibara et<br />

al. 2002) a<br />

In ECM of Betula platyphylla<br />

(Hashimoto and<br />

Hyakumachi 2001)<br />

Mycelium radicis<br />

atrovirens<br />

a Role not fully-established


290 M.C. Brundrett<br />

evolutionary or taxonomic groups, and orchid mycorrhizal associations<br />

are not their primary ecological role (Brundrett 2002). These fungi are<br />

efficient plant colonisers where each has multiple roles as endophytes, parasites,<br />

ECM fungi and saprophytes. We require knowledge about the biology<br />

of these fungi in natural ecosystems to help us understand and control their<br />

plant parasitic activities, as well as their beneficial mycorrhizal associations<br />

with orchids.<br />

16.2.4<br />

Ericoid Mycorrhizal Fungi<br />

Mycorrhizal fungi that associate with members of the Ericaceae (including<br />

the Epacridaceae) include several discrete groups of ascomycetes (McLean<br />

et al. 1999; Monreal et al. 1999; Sharples et al. 2000). Ericoid fungi with<br />

dual roles include Hymenoscyphus ericae, which forms ECM and can also<br />

associate with bryophytes (Table 16.2). However, ericoid fungi also endophytically<br />

colonise roots of ECM hosts without forming mycorrhizas<br />

(Bergero et al. 2000; Piercey et al. 2002). Some strains of the ericoid fungus<br />

Oidiodendron maius, which do not form mycorrhizal associations (Piercey<br />

et al. 2002), appear to be efficient saprophytes (see Chap. 13 by Rice and<br />

Currah). The primary role of ericoid fungi is not clear, as their occurrence<br />

in soils is independent of their host plants [Sharples et al. 2000; Bergero<br />

et al. 2003; see Chaps. 12 (Girlanda et al.) and 14 (Cairney)] and they also<br />

have a high degree of endophytic, or saprophytic competence, or have ECM<br />

associations with trees (Table 16.2).<br />

16.2.5<br />

Endophytic Fungi in Mycorrhizal Roots<br />

Dark septate fungi (called DSF or DSE fungi) are common root endophytes<br />

in many ecosystems – in some cases with dual roles as ECM fungi (Stoyke<br />

and Currah 1991; O’Dell and Trappe 1992; Ahlich and Sieber 1996; see<br />

Chap. 7 by Sieber and Grünig). The DSE root endophytes may provide benefits<br />

to plants (Jumpponen and Trappe 1998). They include Phialocephala<br />

spp. with close relatives that are ECM or ericoid fungi (Vrålstad et al. 2002).<br />

However, DSE fungi do not form mycorrhizal associations as defined by<br />

morphological criteria. Phialocephala fortinii,anECMfungusthatmaynot<br />

benefit host plants, is closely related to other dematiaceous fungi, which<br />

are common root endophytes (Harney et al. 1997). Root endophytes can<br />

act as antagonists of ECM fungi, apparently by competing for space in roots<br />

(Hashimoto and Hyakumachi 2000, 2001). The role of fungi such as DSE<br />

and MRA [Mycelium radicis atrovirens; see Chaps. 7 (Sieber and Grünig)


16 Roles of Endophytic and Mycorrhizal Fungi 291<br />

and 12 (Girlanda et al.)], which commonly share roots with mycorrhizal<br />

fungi, has not been well established.<br />

16.3<br />

Issues with the Identification<br />

and Categorisation of Fungi in Roots<br />

Distinguishing endophytic activity from mycorrhizal associations caused<br />

by the same fungi is often problematic, especially for multifunctional fungi<br />

that are both endophytes and mycorrhizas. The examples discussed above<br />

illustrate how it is essential to use consistent definitions of mycorrhizal<br />

associations based on the structure and development of associations. A key<br />

defining feature of mycorrhizal associations is that they develop in young<br />

roots (Brundrett 2004). Consequently, mycorrhizal formation requires root<br />

growth while endophytic activity is not confined to young healthy roots.<br />

Despitethecommonoccurrenceofendophyticactivitybymycorrhizal<br />

fungi, we have very little knowledge of its ecological significance. Damage<br />

to roots of non-hosts plants caused by attempted colonisation by VAM<br />

and ECM fungi can lead to substantial growth reduction (Allen et al. 1989;<br />

Plattner and Hall 1995; Muthukumar et al. 1997). However, cases of antagonistic<br />

interactions caused by the endophytic activity by mycorrhizal fungi<br />

in non-host plants seem to be rare.<br />

The common occurrence of endophytic fungi in roots may cause confusion<br />

with mycorrhizal fungi, especially when they are identified from DNA.<br />

In most cases, DNA extraction will detect several different fungi within<br />

a root and it may not be clear which are most abundant. Examples of cases<br />

where roles of fungi are uncertain include the study by Allen et al. (2003),<br />

where Sebacina spp. dominated DNA sequences from Gaultheria shallon<br />

(Ericaceae) roots, but did not form ericoid mycorrhizas. Another study<br />

by Bidartondo et al. (2004) found ECM fungi in addition to endophytes<br />

and orchid associates in five orchid species. They concluded that the ECM<br />

fungi were mycorrhizal associates of these orchids, but did not confirm<br />

this by mycorrhizal synthesis. There are numerous other examples in the<br />

recent literature of fungi identified in roots that have been designated as<br />

mycorrhizal without providing sufficient evidence.<br />

Molecular methods used to detect fungi in roots or soils are still relatively<br />

new and more research is required to test the relative efficiency with<br />

which they detect different categories of fungi. In contrast, identification of<br />

mycorrhizalassociationsbymorphologyallowsamuchhigherdegreeof<br />

replication and can be more accurate than DNA-based studies, which sometimes<br />

fail to detect the most important fungi in roots. The interpretation<br />

of fungal presence data provided by molecular tools that allow miniscule


292 M.C. Brundrett<br />

traces of fungi to be detected from almost any substrate will require further<br />

testing of key questions. These questions include: (1) Are the most common<br />

fungi in roots amplified most often? (2) Which of the detected fungi occur<br />

on the surface of roots or are endophytes? (3) How often are traces of DNA<br />

that contaminate samples detected? These questions can be answered by<br />

increasing the replication of sampling strategies, or combining isolation<br />

attempts with DNA extractions (e.g. Allen et al. 2003), but it may be even<br />

better to develop integrated approaches combining molecular and microscopic<br />

techniques. One such study by Sakakibara et al. (2002) found that<br />

microscopic identification of fungi by morphotypes and molecular methods<br />

(PCR-RFLP) were in close agreement, but multiple fungi were obtained<br />

from many mycorrhiza types.<br />

Mycorrhizologists often use the term “endophytes” to refer to both mycorrhizal<br />

and non-mycorrhizal fungi in roots. However, due to the increasing<br />

recognition of the importance of “endophytes” as a specific category of<br />

specialised plant-inhabiting fungi, it is no longer appropriate to call mycorrhizal<br />

fungi endophytes. Types and categories of mycorrhizas are defined<br />

by morphological criteria (see Brundrett 2004), so mycorrhizal fungi need<br />

to be linked to a properly identified mycorrhizal structure. We should designate<br />

fungi as mycorrhizal only if they are isolated from a mycorrhiza by<br />

reliable means and belong to known groups of mycorrhizal associates. It<br />

will be necessary to confirm the mycorrhizal status of fungi by resynthesis,<br />

if it is not clear if they are endophytic or mycorrhizal.<br />

16.4<br />

Evolution of Root-Fungus Associations<br />

An understanding of the capacity for mycorrhizal fungi to grow endophytically<br />

has led to a new theory of how these associations evolved (Brundrett<br />

2002). This theory suggests that the switch to a new mycorrhizal association<br />

type starts with endophytic occupation of roots by fungi, and culminates<br />

in fully functional associations with coordinated development and synchronised<br />

nutrient transfer. The high degree of endophytic competence of<br />

Glomeromycotan fungi apparently has resulted in reacquisition of VAM by<br />

some plant lineages such as the Ericaceae in Hawaii (Koske et al. 1990).<br />

However, the reverse trend, where plant families become nonmycorrhizal<br />

is more common (Brundrett 2002). Other examples of new mycorrhizal<br />

associations, which probably started by endophytic fungal activity, include<br />

plants with dual ECM and VAM associations. These are common in some<br />

plant families that have ancestors with VAM (Brundrett 2002). There also<br />

seem to be intermediate types of ECM associations that involve new lineages<br />

of fungi that are not fully functional, as described above.


16 Roles of Endophytic and Mycorrhizal Fungi 293<br />

The Orchidaceae contain the most examples of plants acquiring new<br />

lineages of symbiotic fungi, in both myco-heterotrophic and green species<br />

(Taylor et al. 2002; Bidartondo et al. 2004). The fungal associates of these<br />

orchids are amazingly diverse, and the only theme that unifies them is that<br />

most are known to be efficient plant colonists as pathogens, endophytes or<br />

ectomycorrhizal associates of other species (Brundrett 2002). This suggests<br />

that most of these fungi were endophytes within orchids before the orchid<br />

evolved means to exploit them for its own purposes. All orchids seem to<br />

require is a fungus that can efficiently invade their roots or stems, but<br />

they seem to have much more success controlling fungi in the Rhizoctonia<br />

alliance than other fungi with similar endophytic competence for reasons<br />

we do not understand.<br />

16.5<br />

Conclusions<br />

As demonstrated by the examples described above, fungi often have several<br />

roles and the primary roles of multifunctional fungi may not be certain.<br />

However,thenumberofrolesfungihavedoesseemtobelimited,presumablybecausetheycannotbeproficientatalloftheseroles.Consequently,<br />

we would expect multifunctional fungi to lose out due to competition from<br />

more efficient fungi that are more highly specialised. There also seem to be<br />

advantages to versatility in fungi. For example, multifunctional fungi seem<br />

to colonise new habitats or substrates more rapidly than specialised fungi,<br />

as shown by studies of mycorrhizal fungal succession after disturbance and<br />

the associations of nursery-grown plants (Jumpponen and Trappe 1998; Yu<br />

et al. 2001). Mycorrhizal fungi take part in a continuum of association types<br />

starting with endophytic associations and concluding with mycorrhizal associations.<br />

However, the endophytic and intermediate roles of these fungi<br />

seem to be less common than mycorrhizal associations, suggesting that<br />

there are clear advantages to fungi from their primary roles with plants.<br />

Recent advances in molecular biology have revealed a much more diverse<br />

and complex picture of the multifunctional nature of mycorrhizal fungi.<br />

Further research is required to determine the relative importance of fungi<br />

detected in roots, as it is not always clear which are endophytes and which<br />

are mycorrhizal associates, or how these roles change with time.<br />

<strong>References</strong><br />

Abdul Karim N (2005) Molecular and enzymic groupings of fungi from tropical orchids of<br />

Western Australia and their patterns of tissue colonisation. PhD Thesis. The University<br />

of Western Australia


294 M.C. Brundrett<br />

Ahlich K, Sieber TN (1996) The profusion of dark septate endophytic fungi in nonectomycorrhizal<br />

fine roots of forest trees and shrubs. New Phytol 132:259–270<br />

Allen MF, Allen EB, Friese CF (1989) Responses of the non-mycotrophic plant Salsola kali<br />

to invasion by vesicular-arbuscular mycorrhizal fungi. New Phytol 111:45–49<br />

Allen TR, Millar T, Berch SM, Berbee ML (2003) Culturing and direct DNA extraction find<br />

different fungi from the same ericoid mycorrhizal roots. New Phytol 160:255–272<br />

Azcón-Aguilar C, Bago B, Barea JM (1999) Saprophytic growth of arbuscular mycorrhizal<br />

fungi. In: Varma A, Hock B (eds) Mycorrhiza, structure, function, molecular biology<br />

and biotechnology. Springer, Berlin Heidelberg New York, pp 391–408<br />

Batty AJ, Dixon KW, Brundrett MC, Sivasithamparam K (2002) Orchid conservation and<br />

mycorrhizal associations. In: Sivasithamparam K, Dixon KW, Barrett RL (eds) Microorganisms<br />

in plant conservation and biodiversity. Kluwer, The Netherlands, pp 195–226<br />

Bayman P, Gonzalez EJ, Fumero, JJ, Tremblay RL (2002) Are fungi necessary? How fungicides<br />

affect growth and survival of the orchid Lepanthes rupestris in the field. J Ecol 90:1002–<br />

1008<br />

Bergero R, Perotto S, Girlanda M, Vidano G, Luppi AM (2000) Ericoid mycorrhizal fungi<br />

are common root associates of a Mediterranean ectomycorrhizal plant (Quercus ilex).<br />

Mol Ecol 9:1639–1649<br />

Bergero R, Girlanda M, Bello F, Luppi AM, Perotto S (2003) Soil Persistence and biodiversity<br />

of ericoid mycorrhizal fungi in the absence of the host plant in a Mediterranean<br />

ecosystem. Mycorrhiza 13:69–75<br />

Bidartondo MI, Burghardt B, Gebauer G, Bruns TD, Read DJ (2004) Changing partners in<br />

the dark: isotopic and molecular evidence of ectomycorrhizal liaisons between forest<br />

orchids and trees. Proc R Soc London B 271:1799–1806<br />

Brundrett MC (1991) Mycorrhizas in natural ecosystems. Adv Ecol Res 21:171–313<br />

Brundrett MC (2002) Coevolution of roots and mycorrhizas of land plants. New Phytol<br />

154:275–304<br />

Brundrett MC (2004) Diversity and classification of mycorrhizal associations. Biol Rev<br />

79:473–495<br />

Brundrett MC, Kendrick WB (1988) The mycorrhizal status, root anatomy, and phenology<br />

of plants in a sugar maple forest. Can J Bot 66:1153–1173<br />

Brundrett M, Bougher N, Dell B, Grove T, Malajczuk N (1996) Working with mycorrhizas<br />

in forestry and agriculture. Australian Centre for International Agricultural Research,<br />

Canberra<br />

Burgess T, Dell B, Malajczuk N (1994) Variations in mycorrhizal development and growth<br />

stimulation by 20 Pisolithus isolates inoculated on to Eucalyptus grandis W. Hill ex<br />

Maiden. New Phytol 127:731–739<br />

Cázares E, Trappe JM (1993) Vesicular endophytes in roots of the Pinaceae. Mycorrhiza<br />

2:153–56<br />

Chambers SM, Williams PG, Seppelt RD, Cairney JWG (1999) Molecular identification<br />

of Hymenoscyphus sp. from rhizoids of the leafy liverwort Cephaloziella exiliflora in<br />

Australia and Antarctica. Mycol Res 103:286–288<br />

Chen YL, Dell B, Brundrett MC (2000) Effects of ectomycorrhizas and vesicular-arbuscular<br />

mycorrhizas, alone or in competition, on root colonization and growth of Eucalyptus<br />

globulus and E. urophylla. New Phytol 146:545–556<br />

Chilvers GA, Gust LW (1982) Comparison between the growth rates of mycorrhizas, uninfected<br />

roots and a mycorrhizal fungus of Eucalyptus st-johnii R. T. Bak. New Phytol<br />

91:453–66<br />

Cordier C, Pozo MJ, Barea JM, Gianinazzi S, Gianinazzi-Pearson V (1998) Cell defence<br />

responses associated with localized and systemic resistance to Phytophthora parasitica


16 Roles of Endophytic and Mycorrhizal Fungi 295<br />

induced in tomato by an arbuscular mycorrhizal fungus. Mol Plant-Microb Interact<br />

11:1017–1028<br />

Cornwell WK, Bedford BL, Chapin CT (2001) Occurrence of arbuscular mycorrhizal fungi<br />

in a phosphorus-poor wetland and mycorrhizal responses to phosphorus fertilization.<br />

Am J Bot 88:1824–1829<br />

Currah RS, Zelmer CD, Hambleton S, Richardson KA (1997) Fungi from orchid mycorrhizas.<br />

In: Arditti J, Pridgeon AM (eds) Orchid biology: reviews and perspectives, VII. Kluwer,<br />

Dordrecht, pp 117–170<br />

Dahlstrom JL, Smith JE, Weber NS (2000) Mycorrhiza-like interaction by Morchella with<br />

species of the Pinaceae in pure culture synthesis. Mycorrhiza 9:279–285<br />

Demchenko K, Winzer T, Stougaard J, Parniske M, Pawlowski K (2004) Distinct roles of<br />

Lotus japonicus SYMRK and SYM15 in root colonisation and arbuscular formation.<br />

New Phytol 163:381–392<br />

Duckett JG, Read DJ (1995) Ericoid mycorrhizas and rhizoid-ascomycete associations in<br />

liverworts share the same mycobiont: isolation of the partners and resynthesis of the<br />

associations in vitro. New Phytol 129:439–447<br />

Eason WR, Newman EI, Chuba PN (1991) Specificity of interplant cycling of phosphorus:<br />

the role of mycorrhizas. Plant Soil 137:267–274<br />

Gill WM, Lapeyrie F, Gomi T, Suzuki K (1999) Tricholoma matsutake – an assessment of in<br />

situ and in vitro infection by observing cleared and stained roots. Mycorrhiza 9:227–231<br />

Giovannetti M, Sbrana C (1998) Meeting a non-host: the behaviour of AM fungi. Mycorrhiza<br />

8:123–130<br />

Glen M, Tommerup IC, Bougher NL, O’Brien PA (2002) Are Sebacinaceae common and<br />

widespread ectomycorrhizal associates of Eucalyptus species in Australian forests. Mycorrhiza<br />

12:243–247<br />

Hall IR (1976) Vesicular mycorrhizas in the orchid Corybas macranthus.TransBrMycolSoc<br />

66:160<br />

Harley JL, Harley EL (1987) A check-list of mycorrhiza in the British flora. New Phytol<br />

105[Suppl 2]:1–102<br />

Harney SK, Rogers SO, Wang CJK (1997) Molecular characterization of dematiaceous root<br />

endophytes. Mycol Res 101:1397–1404<br />

Harrington TJ, Mitchell DT (2002) Colonization of root systems of Carex flacca and C. pilulifera<br />

by Cortinarius (Dermocybe) cinnamomeus. Mycol Res 106:452–459<br />

Hashimoto Y, Hyakumachi M (2000) Quantities and types of ectomycorrhizal and endophytic<br />

fungi associated with Betula platyphylla var. japonica seedlings during the initial<br />

stage of establishment of vegetation after disturbance. Ecol Res 15:21–31<br />

Hashimoto Y, Hyakumachi M (2001) Effects of isolates of ectomycorrhizal fungi and endophytic<br />

Mycelium radicis atrovirens that were dominant in soil from disturbed sites on<br />

growth of Betula platyphylla var. japonica seedlings. Ecol Res 16:117–125<br />

Imhof S (2001) Subterranean structures and mycotrophy of the achlorophyllous Dictyostega<br />

orobanchoides (Burmanniaceae). Rev Biol Trop 49:239–247<br />

Joner EJ, Jakobsen I (1995) Growth and extracellular phosphate activity of arbuscular<br />

mycorrhizal hyphae as influenced by soil organic matter. Soil Biol Biochem 9:1153–<br />

1159<br />

Jumpponen A, Trappe JM (1998) Dark septate endophytes: a review of facultative biotrophic<br />

root-colonizing fungi. New Phytol 140:295–310<br />

KõljalgU,DahlbergA,TaylorAFS,LarssonE,HallenbergN,StenlidJ,LarssonK-H,Fransson<br />

PM, Kårén O, Jonsson L (2000) Diversity and abundance of resupinate thelephoroid<br />

fungi as ectomycorrhizal symbionts in Swedish boreal forests. Mol Ecol 9:1985–1996<br />

Koske RE (1984) Spores of VAM fungi inside spores of VAM fungi. Mycologia 76:853–862


296 M.C. Brundrett<br />

Koske RE, Gemma JN, Englander L (1990) Vesicular-arbuscular mycorrhizae in Hawaiian<br />

Ericales. Am J Bot 77:64–68<br />

Kottke I, Beiter A, Weiss M, Haug I, Oberwinkler F, Nebel M (2003) Heterobasidiomycetes<br />

form symbiotic associations with hepatics: Jungermanniales have sebacinoid mycobionts<br />

while Aneura pinguis (Metzgeriales) is associated with a Tulasnella species.<br />

Mycol Res 107:957–968<br />

Kristiansen KA, Taylor DL, Kjøller R, Rasmussen HN, Rosendahl S (2001) Identification<br />

of mycorrhizal fungi from single pelotons of Dactylorhiza majalis (Orchidaceae) using<br />

single-strand conformation polymorphism and mitochondrial ribosomal large subunit<br />

DNA sequences. Mol Ecol 10:2089–2093<br />

Kuldau GA, Yates IE (2000) Evidence for Fusarium endophytes in cultivated and wild plants.<br />

In: Bacon CW, White JF Jr (eds) Microbial endophytes. Decker, New York, pp 85–117<br />

Malajczuk N, Dell B, Bougher NL (1987) Ectomycorrhiza formation in Eucalyptus III.<br />

Superficial ectomycorrhizas initiated by Hysterangium and Cortinarius species. New<br />

Phytol 105:421–428<br />

McKormick MK, Whigham DF, O’Neill (2004) Mycorrhizal diversity of photosynthetic terrestrial<br />

orchids. New Phytol 163:425–438<br />

McLean CB, Cunnington JH, Lawrie AC (1999) Molecular diversity within and between<br />

ericoid endophytes from the Ericaceae and Epacridaceae. New Phytol 144:351–358<br />

Monreal M, Berch SM, Berbee M (1999) Molecular diversity of ericoid mycorrhizal fungi.<br />

Can J Bot 77:1580–1594<br />

Muthukumar T, Udaiyan K, Karthikeyan A, Manian S (1997) Influence of native endomycorrhiza,<br />

soil flooding and nurse plant on mycorrhizal status and growth of purple<br />

nutsedge (Cyperus rotundus L.). Agric Ecosyst Environ 61:51–58<br />

Muthukumar T, Udaiyan K, Shanmughavel P (2004) Mycorrhiza in sedges – an overview.<br />

Mycorrhiza 14:65–77<br />

Mursidawati S (2003) Mycorrhizal association, propagation and conservation of the mycoheterotrophic<br />

orchid Rhizanthella gardneri. MSc.Thesis,TheUniversityofWestern<br />

Australia<br />

Newsham KK, Fitter AH, Watkinson AR (1995) Arbuscular mycorrhiza protect an annual<br />

grass from root pathogenic fungi in the field. J Ecol 83:991–1000<br />

Ocampo JA (1986) Vesicular-arbuscular mycorrhizal infection of “host” and “non-host”<br />

plants: effect on the growth responses of the plants and competition between them. Soil<br />

Biol Biochem 18:607–610<br />

O’Dell T, Trappe JM (1992) Root endophytes of lupin and some other legumes of Northwestern<br />

U.S.A. New Phytol 122:479–485<br />

Otero JT, Ackerman JD, Bayman P (2002) Diversity and host specificity of endophytic<br />

Rhizoctonia-like fungi from tropical orchids. Am J Bot 89:1852–1858<br />

Pamphile JA, Azevedo JL (2002) Molecular characterization of endophytic strains of Fusarium<br />

verticillioides (= Fusarium moniliforme) from maize (Zea mays L). World J Microbiol<br />

Biotechnol 18:391–396<br />

Peterson RL, Massicotte HB (2004) Exploring structural definitions of mycorrhizas, with<br />

emphasis on nutrient-exchange interfaces. Can J Bot 82:1074–1088<br />

Piercey MM, Thormann MN, Currah RS (2002) Saprobic characteristics of three fungal<br />

taxa from ericalean roots and their associations with the roots of Rhododendron groenlandicum<br />

and Picea mariana in culture. Mycorrhiza 12:175–180<br />

Plattner I, Hall IR (1995) Parasitism of non-host plants by the mycorrhizal fungus Tuber<br />

melanosporum. Mycol Res 99:1367–1370<br />

Powell CL (1975) Rushes and sedges are non-mycotrophic. Plant Soil 42:481–484


16 Roles of Endophytic and Mycorrhizal Fungi 297<br />

Rabatin SC, Rhodes LH (1982) Acaulospora bireticulata inside orbatid mites. Mycologia<br />

74:859–861<br />

Redman RS, Dunigan DD, Rodriguez RJ (2001) Fungal symbiosis from mutualism to parasitism:<br />

who controls the outcome, host or invader? New Phytol 151:705–716<br />

Richardson KA, Currah RS (1995) The fungal community associated with the roots of some<br />

rainforest epiphytes of Costa Rica. Selbyana 16:49–73<br />

Saikkonen K, Faeth SH, Helander M, Sullivan TJ (1998) Fungal endophytes: a continuum of<br />

interactions with plants. Annu Rev Ecol Syst 29:319–343<br />

Sakakibara SM, Jones MD, Gillespie M, Hagerman SM, Forrest ME, Simard SW, Durall DM<br />

(2002) A comparison of ectomycorrhiza identification based on morphotyping and<br />

PCR-RFLP analysis. Mycol Res 106:868–878<br />

Selosse MA, Bauer R, Moyersoen B (2002a) Basal hymenomycetes belonging to the Sebacinaceae<br />

are ectomycorrhizal on temperate deciduous trees. New Phytol 155:183–195<br />

Selosse M-A, Weiss M, Jany J-L, Tillier A (2002b) Communities and populations of sebacinoid<br />

basidiomycetes associated with the achlorophyllous orchid Neottia nidus-avis (L.)<br />

L.C.M. Rich. and neighbouring tree ectomycorrhizae. Mol Ecol 11:1831–1844<br />

Sen R, Hietala AM, Zelmer CD (1999) Common anastomosis and internal transcribed<br />

spacer RFLP groupings in binucleate Rhizoctonia isolates representing root endophytes<br />

of Pinus sylvestris, Ceratorhiza spp. from orchid mycorrhizas and a phytopathogenic<br />

anastomosis group. New Phytol 144:331–341<br />

Sharples JM, Chambers SM, Meharg AA, Cairney JWG (2000) Genetic diversity of rootassociated<br />

fungal endophytes from Calluna vulgaris at contrasting field sites. New<br />

Phytol 148:153–162<br />

Sivasithamparam K (1998) Root cortex – the final frontier for biocontrol of root-rot<br />

with fungal antagonists: a case study on a sterile red fungus. Annu Rev Phytopathol<br />

36:439–452<br />

Smith JE, Johnson KA, Cázares E (1998) Vesicular mycorrhizal colonisation of seedlings of<br />

Pinaceae and Betulaceae after spore inoculation with Glomus intraradices. Mycorrhiza<br />

7:279–285<br />

St John TV, Coleman DC, Reid CPP (1983) Growth and spatial distribution of nutrientabsorbing<br />

organs: selective exploitation of soil heterogeneity. Plant Soil 71:487–493<br />

Stoyke G, Currah RS (1991) Endophytic fungi from the mycorrhizae of alpine ericoid plants.<br />

Can J Bot 69:347–352<br />

Taylor DL, Bruns TD, Leake JR, Read DJ (2002) Mycorrhizal specificity and function in<br />

myco-heterotrophic plants. In: Van der Heijden MGA, Sanders IR (eds) The ecology of<br />

mycorrhizas. Springer, Berlin Heidelberg New York, pp 375–414<br />

Tester M, Smith SE, Smith FA (1987) The phenomenon of “nonmycorrhizal” plants. Can<br />

J Bot 65:419–431<br />

Urban A, Weiss M, Bauer R (2003) Ectomycorrhizas involving sebacinoid mycobionts. Mycol<br />

Res 107:3–14<br />

Van der Heijden EW (2001) Differential benefits of arbuscular mycorrhizal and ectomycorrhizal<br />

infection of Salix repens. Mycorrhiza 10:185–193<br />

Vrålstad T, Fossheim T, Schumacher T (2000) Piceirhiza bicolorata – the ectomycorrhizal<br />

expression of the Hymenoscyphus ericae aggregate? New Phytol 145:549–563<br />

Vrålstad T, Schumacher T, Taylor FS (2002) Mycorrhizal synthesis between fungal strains of<br />

the Hymenoscyphus ericae aggregate and potential ectomycorrhizal and ericoid hosts.<br />

New Phytol 153:143–152<br />

Vujanovic V, St-Arnaud M, Barabé D, Thibeaults G (2000) Viability testing of orchid seed<br />

and the promotion of colouration and germination. Ann Bot 86:79–86


298 M.C. Brundrett<br />

Warcup JH (1981) The mycorrhizal relationships of Australian orchids. New Phytol<br />

87:371–38<br />

Warcup JH (1985) Rhizanthella gardneri (Orchidaceae), its Rhizoctonia endophyte and close<br />

association with Melaleuca uncinata (Myrtaceae) in Western Australia. New Phytol<br />

99:273–280<br />

Warner A (1984) Colonisation of organic matter by mycorrhizal fungi. Trans Br Mycol Soc<br />

82:352–354<br />

Yu TEJ-C, Egger KN, Peterson RL (2001) Ectendomycorrhizal associations – characteristics<br />

and functions. Mycorrhiza 11:167–177


17<br />

17.1<br />

Introduction<br />

Isolation Procedures<br />

for Endophytic Microorganisms<br />

Johannes Hallmann, Gabriele Berg, Barbara Schulz<br />

There are good reasons for isolating endophytic microorganisms, e.g. for<br />

their characterisation, for studying population dynamics and diversity, use<br />

of microbial inoculants to improve plant growth and plant health, and<br />

as sources of novel biologically active secondary metabolites (Schulz et<br />

al. 2002; Strobel 2003; Schulz and Boyle 2005). The isolation procedure<br />

is a critical and important step in working with endophytic bacteria and<br />

fungi. It should be sensitive enough to recover endophytic microorganisms,<br />

butatthesametimebestrongenoughtoeliminateepiphytesfromtheroot<br />

surface. In practice, this is often difficult, because microorganisms attach to<br />

plant cells/surfaces or hide in intercellular niches, thus evading the isolation<br />

procedure. Besides, there is no sharp delineation between the external and<br />

internal plant environment. Some bacteria constantly move between these<br />

two microhabitats, e.g. fungal mycelia grow from the root surface into the<br />

roots. Therefore, in order to be effective, the isolation procedure must<br />

be adapted to the respective tissues and microorganisms. Some isolation<br />

procedures are suitable only for certain plant tissues, whereas others favour<br />

local or systemic colonisers.<br />

Themostcommonlyusedisolationprocedurescombinesurfacesterilisation<br />

of the root tissue with either maceration of the plant tissue and streaking<br />

onto nutrient agar, or plating small sterilised segments onto nutrient<br />

agar. Methods that avoid surface sterilisation are also available, e.g. vacuum<br />

or pressure extraction. In the past, isolation procedures focused primarily<br />

on culturable microorganisms. However, increasing interest in nonculturable<br />

endophytic microorganisms has recently led to the application<br />

Johannes Hallmann: Federal Biological Research Centre for Agriculture and Forestry, Institute<br />

for Nematology and Vertebrate Research, Toppheideweg 88, 48161 Münster, Germany,<br />

E-mail: j.hallmann@bba.de<br />

Gabriele Berg: Technical University of Graz, Department of Environmental Biotechnology,<br />

Petersgasse 12, 8010 Graz, Austria<br />

Barbara Schulz: Institute of Microbiology, Technical University of Braunschweig, Spielmannstraße<br />

7, 38106 Braunschweig, Germany<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


300 J. Hallmann et al.<br />

of molecular methods for their identification. This chapter summarises isolation<br />

procedures for culturable and detection methods for non-culturable<br />

endophytic bacteria and fungi of plant roots and discusses the strengths<br />

and weaknesses of each procedure. Examples are given on how the choice<br />

of the procedure can affect the microbial spectrum being isolated. The information<br />

provided here should enable the researcher to specifically select<br />

the method most suitable to meet their research objectives. The reader is<br />

also referred to the review by Sieber (2002), who summarised the available<br />

sterilisation and isolation procedures for fungi in plant roots.<br />

17.2<br />

Surface Sterilisation<br />

Theoretically, the sterilising agent should kill any microbe on the plant<br />

surface without affecting the host tissue and the endophytic microorganisms.<br />

However, this is difficult to achieve because in time the agent may<br />

penetrate the root tissue. Conditions required to kill the last microbe on<br />

the surface may already be lethal for some endophytic microorganisms. In<br />

general, surface sterilisation of root tissue consists of the following steps:<br />

(1) thorough washing of the root tissue under tap water to remove adhering<br />

soil particles and the majority of microbial surface epiphytes and incidentals,<br />

(2) pre-treatment (optional) to eliminate hydrophobic substances on<br />

the plant surface and to provide better access for the sterilising agent to<br />

the root surface, (3) surface sterilisation to eliminate remaining microbial<br />

colonisers from the root surface, (4) several rinses under aseptic conditions<br />

in a sterile washing solution and (5) sterility check to confirm complete<br />

sterilisation of the root surface. It is very important that sterility is guaranteed<br />

for all tools and during all steps of this procedure, and that the<br />

researcher optimises the procedure for each plant tissue, since sensitivity<br />

varies with species, age and surface properties.<br />

All steps in the sterilisation procedure should be conducted with sterilised<br />

tools, e.g. pincers, and preferably under a laminar flow hood. Between<br />

steps the tissue should be blotted on filter paper to avoid dilution of the<br />

sterilising agents.<br />

17.2.1<br />

Pre-treatment<br />

Roots usually do not require special pre-treatment, since their surfaces do<br />

not contain hydrophobic substances like, e.g., the waxes of leaves. Washing<br />

with tap water and/or soft brushing is usually adequate. However, sonification<br />

may be used to dislodge soil and organic matter from the roots prior


17 Isolation Procedures for Endophytic Microorganisms 301<br />

to surface sterilisation (Holdenrieder and Sieber 1992; Coombs and Franco<br />

2003a).<br />

17.2.2<br />

Sterilising Agents<br />

Commonly used sterilising agents are sodium hypochlorite (Gardner et al.<br />

1982; Quadt-Hallmann et al. 1997; Schulz et al. 1993; Sieber 2002), ethanol<br />

(Dong et al. 1994; Gagné et al. 1987), and hydrogen peroxide (McInroy<br />

and Kloepper 1994; Misahgi and Donndelinger 1990; Sieber 2002). For<br />

health reasons, mercuric chloride should be used only very carefully and<br />

exclusively for plant tissues that cannot otherwise be sterilised (Gagné et al.<br />

1987; Hollis 1951; O’Dell and Trappe 1992; Sriskandarajah et al. 1993). Note<br />

that sodium hypochlorite, an excellent oxidant, decomposes spontaneously<br />

in storage. Other, less frequently used, agents include propylene oxide<br />

vapour (Sardi et al. 1992) and formaldehyde (Schulz et al. 1993; Cao et<br />

al. 2002). Woody root tissues may also be dipped for 15 s in 95% ethanol<br />

and flame sterilised as employed for alfalfa roots by Gagné et al. (1987). In<br />

addition, the outer layer of woody root tissue can be aseptically peeled and<br />

the internal plant tissue can be excised using a sterile cork borer (Maifeld<br />

1998; Reiter et al. 2002; Zinniel et al. 2002).<br />

Combinations of agents can improve the effectiveness of surface sterilisation.<br />

For example, a combination of physical and chemical sterilisation has<br />

given good results with lignified roots (Table 17.1; Görke 1998). A common<br />

protocol involves a three-step procedure with ethanol, sodium hypochlorite,<br />

and then ethanol again (Schulz et al. 1993; Bills 1996; Sieber 2002).<br />

For example, for wheat roots, Coombs and Franco (2003b) applied 99%<br />

ethanol for 1 min, 3.1% sodium hypochlorite for 6 min and 99% ethanol<br />

for 30 sec. Sieber (2002) suggested omitting the initial soak in ethanol because<br />

it leads to contraction of the plant tissues, perhaps entrapping some<br />

epiphytic conidia, spores or mycelia. All the conidia of Penicillium sp. on<br />

roots of Norway spruce (artificially contaminated) were killed when surface<br />

sterilisation with hydrogen peroxide was used without the initial soak in<br />

ethanol (Sieber 2002). Examples of commonly used sterilising agents and<br />

conditions for their application are given in Table 17.1.<br />

In general, a more stringent sterilisation procedure can be used with<br />

older and lignified plant tissue. Incubation times as well as concentrations<br />

of the sterilising agents directly affect the results. For example, an increase<br />

in the concentration of sodium hypochlorite from 3 to 6% active chlorine<br />

under otherwise constant conditions increases the number of root samples<br />

with complete sterilisation by 40%, but at the same time decreases<br />

the population densities of endophytic bacteria from 4.36×10 3 cfu/ml to


302 J. Hallmann et al.<br />

Table 17.1. Protocols for surface sterilisation of root tissue<br />

Procedures and sterilising agents Incubation Root type, age and/or Plant Reference<br />

time diameter<br />

Bacteria<br />

a) Wash under running tap water 5- to 7-week-old roots Cucumis sativus, Gossyium Hallmann et al. 1998;<br />

b) 1–3.1% Sodium hypochlorite hirsutum, Pinus contorta McInroy and Kloepper 1995;<br />

var. latifolia, Zea mays Shishido et al. 1995<br />

a 1–2 min<br />

c)Twotofourrinsesinsterilewater<br />

a) 99% Ethanol 1 min 1- to 3-month-old roots Triticum aestivum Coombs and Franco 2003a<br />

b) 3.1% Sodium hypochlorite 6 min<br />

c) 99% Ethanol 0.5 min<br />

a) Wash with soapy water 4-to 28-month-old roots Medicago sativa Gagné et al. 1987<br />

b) Wash thoroughly with tap water<br />

c) 0.2 M HgCl2 in 50% ethanol 4 min<br />

d) Three rinses in sterile water 1 min<br />

a) Wash under running tap water 3 min 1- to 4-year-old roots Citrus jambhiri Gardner et al. 1982<br />

b) Dip in 95% ethanol<br />

c) Flame<br />

Sardi et al. 1992<br />

Various herbaceous<br />

and woody plants<br />

d)Rinseinsterilewater<br />

a) Wash under running tap water Roots 1- to 5 mm in<br />

b) Exposure to propylene oxide vapor 1 h diameter<br />

20 min 13- to 14-week-old roots Solanum tuberosum Sessitsch et al. 2002<br />

Physical sterilisation<br />

a) Shake vigorously in 0.9% NaCl solution<br />

containing 0.3 g acid-washed glass beads<br />

b) Five Rinses in sterile water<br />

Fungi<br />

36% Formaldehyde Musa acuminata Cao et al. 2002


17 Isolation Procedures for Endophytic Microorganisms 303<br />

Table 17.1. (continued)<br />

Plant Reference<br />

Oberholzer-Tschütscher 1982;<br />

Sieber et al. 1988;<br />

Crous et al. 1995b Triticum aestivum,<br />

Erica carnea<br />

Skipp and Christensen 1989;<br />

Kattner and Schönhar 1990;<br />

Hambleton and Currah 1997;<br />

Stoyke and Currah 1991b .<br />

Hallmann and Sikora 1994<br />

Lolium perenne, Picia abies<br />

various species of Ericaceae,<br />

various alpine plant species<br />

Lycopersicon esculentum<br />

Procedures and sterilising agents Incubation Root type, age and/or<br />

time diameter<br />

a) Wash under running tap water Nodule roots, fine hair<br />

b) 96% Ethanol 0.5–1.0 min roots, nonlignified<br />

c) 2–2.5% Sodium hypochlorite 1–4 min roots, lignified roots<br />

d) 96% Ethanol 0.5 min<br />

a) Wash under running tap water Adventitious roots, fine<br />

b) 0.3–2% Sodium hypochlorite 1–3 min hair roots<br />

c)Rinseinsterilewater<br />

Boyle et al. 2001<br />

Petrini et al. 1992;<br />

Werner et al. 1997b O’Dell and Trappe 1992 b<br />

a) Wash under running tap water Roots Hordeum vulgare,<br />

b) 70% Ethanol 0.5 min<br />

Phaseolus vulgarum<br />

c) 1–3% Sodium hypochlorite 1–3 min<br />

d) 70% Ethanol 0.5 min<br />

a) Wash under running tap water 1 min Rhizomes;<br />

Pteridium aquilinum;<br />

b) 75% Ethanol 3 min 6- to 9-month-old roots Aphelandra tetragona<br />

c) 3–5% Sodium hypochloritea 0.5 min<br />

d) 75% Ethanol<br />

0.01% Mercuric chloride 5–15 min Fine roots Lupinus spp.,<br />

Oxytropis campestris<br />

O’Dell and Trappe 1992 b<br />

5% Hydrogen peroxide 5–15 min Fine roots Lupinus spp.,<br />

Oxytropis campestris


304 J. Hallmann et al.<br />

Table 17.1. (continued)<br />

Procedures and sterilising agents Incubation Root type, age and/or Plant Reference<br />

time diameter<br />

a) Wash under running water Adventitious roots Oryza sativa Fisher and Petrini 1992b b) 75% Ethanol 1 min<br />

c) 20% Sodium hypochlorite 3 min<br />

d) 75% Ethanol 0.5 min<br />

a) Wash under running water Non-ectomycorrhizal Abies alba Ahlich and Sieber 1996<br />

roots, 0–5.3 mm in<br />

diameter<br />

b<br />

b) 99% Ethanol 1 min<br />

c) 35% Hydrogen peroxide 5 min<br />

d) 99% Ethanol 0.5 min<br />

Picea abies Maifeld 1998 b<br />

Görke 1998 b<br />

Fagus sylvatica, Picea abies,<br />

Pinus sylvestris, Betula<br />

pendula<br />

Physical sterilisation<br />

a) Cut core into pieces Boring cores from which<br />

b) Move pieces quickly through<br />

bark has been removed<br />

flame of Bunsen burner<br />

Physical + chemical sterilisation<br />

a) Wash under running tap water 2- to 9-year-old roots<br />

b) 70% Ethanol 1 min undergoing secondary<br />

c) 10% Sodium hypochlorite 3 min growth<br />

d) 70% Ethanol 0.5 min<br />

e) Rinse twice in sterile water<br />

f) Discard bark<br />

g) Move pieces quickly through<br />

flame of Bunsen burner<br />

aConcentrations of sodium hypochlorite represent percent active chlorine<br />

bAdapted from Sieber 2002


17 Isolation Procedures for Endophytic Microorganisms 305<br />

5.75 × 10cfu/ml (A. Munif, personal communication). The effect of the<br />

sterilising agent penetrating the plant tissue on the bacterial cells was visually<br />

demonstrated by treating the root tissue with a tetrazolium-phosphate<br />

buffer solution (Patriquin and Döbereiner 1978). While metabolically active<br />

bacterial cells reduced tetrazolium and became stained, cells killed<br />

by the sterilisation remain unstained. Alternatively, apoplastic dyes such<br />

as 3-hydroxy-5,8,10-pyrenetrisulfonate (PTS) or 4.4 ′ -bis (2-sulfostyryl) biphenylcanbeusedtomonitortherelativeprogressionoftheappliedagent<br />

into the plant tissue (Petersen et al. 1981).<br />

17.2.3<br />

Surfactants<br />

Surfactants such as Tween 20 (Mahaffee and Kloepper 1997), Tween 80<br />

(Sturz 1995) or Triton X-100 (Misaghi and Donndelinger 1990) added to<br />

the sterilising agent reduce the surface tension and allow the agent to reach<br />

into niches and grooves beyond the epidermal cells (Bills 1996; Hallmann<br />

et al. 1997a).<br />

17.2.4<br />

Rinsing<br />

After each treatment, the plant tissue should be rinsed repeatedly in sterile<br />

water.<br />

17.2.5<br />

Sterility Check and Optimisation<br />

Only if complete surface sterilisation of the root tissue is confirmed, can the<br />

isolated microorganisms be assumed to be endophytes. Validation of the<br />

surface sterilisation procedure can be done by (1) imprinting the surfacesterilised<br />

plant tissue onto nutrient media (Pleban et al. 1995; Shishido et<br />

al. 1995; Schulz et al. 1998), (2) culturing aliquots of water from the last<br />

rinsing onto nutrient media (McInroy and Kloepper 1994) or (3) dipping<br />

the roots into nutrient broth (Gagné et al. 1987). All three methods gave<br />

comparable results (Musson et al. 1995). If no microbial growth occurs on<br />

themedium,surfacesterilisationisconsideredcomplete.Afurthercheckis<br />

to test the effect of the sterilising agent directly on fungal or bacterial cells.<br />

Rootsaredippedintoafungalorbacterialsuspensionofknowndensity,<br />

slightly dried, surface sterilised and subjected to a sterility check (Petrini<br />

1984; Coombs and Franco 2003a).


306 J. Hallmann et al.<br />

17.3<br />

Culture of Tissue and Plant Fluid of Sterilised Roots<br />

on Nutrient Medium<br />

17.3.1<br />

Segments<br />

The surface-sterilised root is cut aseptically into segments, which are plated,<br />

i.e. pressed directly, onto an appropriate nutrient medium (see below). It<br />

is very important to cut the tissue into very small and thin segments, since<br />

colonisation may be very limited (Carroll 1995; Boyle et al. 2001). For fungal<br />

isolation, the nutrient medium should be supplemented with antibiotics to<br />

suppress bacterial growth. When isolating fungi, antifungal substances may<br />

also be added to retard the growth of fast-growing fungi; when isolating<br />

bacteria to inhibit growth of fungi (see below). Endophytic microorganisms<br />

that emerge from the root fragments are subsequently transferred to<br />

freshmedia.Thismethodiscommonlyusedforendophyticfungi(Schulz<br />

et al. 1993; Hallmann and Sikora 1994; Bills 1996) and actinobacteria (Sardi<br />

et al. 1992; Coombs and Franco 2003a), but is rarely appropriate for endophytic<br />

bacteria (Araújo et al. 2001). This method selects for fast-growing<br />

microorganisms, therefore representing more a qualitative than a quantitative<br />

approach. When studying microbial diversity, slow growing microorganisms<br />

may be under-represented. In addition, concomitant growth<br />

of two or more microorganisms necessitates subsequent separation of the<br />

isolates and may also result in a lower number of colonies being recovered<br />

(Elvira-Recuenco and van Vuurde 2000).<br />

17.3.2<br />

Maceration of Root Tissue<br />

Maceration is the preferred method for the isolation of endophytic bacteria<br />

from surface sterilised tissue, but may also be employed for isolating endophytic<br />

fungi (Sieber 2002), particularly slow growing ones. Theoretically,<br />

it captures all endophytic colonisers of the root tissue, i.e. colonisers of<br />

the root cortex as well as of the vascular tissue, systemic as well as local<br />

colonisers, and intercellular as well as intracellular colonisers. Maceration<br />

is useful for determining the broad spectrum of culturable bacterial and<br />

fungal endophytes; however, it excludes obligate biotrophs. To improve<br />

maceration and form a homogenous suspension, sterile water or sterile<br />

buffer solution is usually added to the surface-sterilised root tissue prior to<br />

maceration. Maceration of the surface-sterilised tissue can be performed


17 Isolation Procedures for Endophytic Microorganisms 307<br />

with mortar and pestle or with mechanical devices such as a Klecco tissue<br />

pulveriser (Mahaffee and Kloepper 1997), a Polytron homogeniser (Zinniel<br />

et al. 2002) or a blender (Araújo et al. 2001), depending on sample size<br />

and the hardness of the plant material. Especially for woody plant tissues,<br />

processing by hand can be laborious and exhausting and mechanical devices<br />

are advisable. For the latter, optimal time and intensity of maceration<br />

need to be empirically determined. Following maceration of the surfacesterilised<br />

root tissue, the suspension is streaked onto nutrient medium.<br />

Sterility has to be guaranteed during all steps of this procedure. Since<br />

plant enzymes and toxins released during maceration can inactivate or<br />

kill endophytic microorganisms, temperature should not increase to lethal<br />

levels,necessitatingcooling;thesampleshouldbeimmediatelydilutedto<br />

decrease the concentrations of toxic compounds to ineffective levels. Alternatively,<br />

substances can be added to buffer the toxic compounds, e.g.<br />

polyvinylpyrrolidone (PVP) or EDTA. To minimise the time required for<br />

the entire procedure from surface sterilisation to maceration, Musson et al.<br />

(1995) described a microtitre plate method in which all steps are included<br />

in one plate. This method resulted in a higher detection limit and improved<br />

sterility, but was limited to small sample sizes.<br />

17.3.3<br />

Centrifugation of Root Tissue<br />

Centrifugation is commonly used to collect the intercellular (apoplastic)<br />

fluid of plant tissue (De Wit and Spikman 1982; Boyle et al. 2001), but<br />

may also be used to extract endophytic microorganisms. This technique<br />

has been successfully applied for the isolation of endophytic bacteria from<br />

sugarcane stems (Dong et al. 1994), but supposedly is also suitable for the<br />

isolation of endophytic microorganisms from root tissue. Depending on<br />

sample size, sterile Eppendorf tubes or larger glass test tubes are used for<br />

centrifugation of the surface-sterilised tissues. Most of the apoplastic fluid<br />

will be removed at 3,000g (Dong et al. 1994). The collected plant fluid is<br />

then streaked on nutrient medium for bacterial and fungal recovery. This<br />

approach avoids maceration of the root tissue, as demonstrated by cryoscanning<br />

electron microscopy, which showed that the cells were still intact<br />

and thus no contamination with symplastic fluid had occurred (Dong et al.<br />

1994). Although this technique seems to be time-consuming, requiring both<br />

surface sterilisation and centrifugation, several samples can be centrifuged<br />

atthesametimeandtheoveralltimerequirementmayineffectbenomore<br />

than for other methods.<br />

Whereas surface sterilisation, followed by subsequent plating of macerated<br />

tissue segments or of centrifuged fluid, is a simple and valuable method


308 J. Hallmann et al.<br />

that produces consistent results, it has its limitations: (1) it is laborious; (2)<br />

microorganismscanhideinnichesthatarenotreachedbythesterilising<br />

agent or “stick” to the host tissue, resulting in false identifications of endophytes;<br />

and (3) the sterilising agent may penetrate the root tissue and kill<br />

endophytes, so that microbial densities will be underestimated. These disadvantages<br />

can be avoided by using techniques to isolate microorganisms<br />

that do not involve surface sterilisation.<br />

17.4<br />

Vacuum and Pressure Extraction<br />

To bypass the above mentioned problems, in particular involving surface<br />

sterilisation, alternative methods have been developed. These techniques<br />

use vacuum or pressure to collect plant fluid from the root tissue. Both<br />

approaches have in common that plant fluid of the conducting elements<br />

and adjacent intercellular spaces is collected, as discussed for centrifugation,<br />

thus reaching two niches often considered favourable for systemic<br />

colonisers (Hallmann et al. 1997a, 1997b). However, it does not isolate<br />

those endophytes that grow intracellularly. Bell et al. (1995) and Gardner et<br />

al. (1982) used an aseptic vacuum extraction technique to isolate bacteria<br />

from the xylem of grapevine and citrus roots. Cohen (1999) described an<br />

oversized vacuum filtration unit for large scale isolation of fungi from leaf<br />

tissue, which might also be adaptable for root tissue. The Scholander pressurebombcommonlyusedformeasuringplantwaterstatusisalsoeffective<br />

for the isolation of endophytic microorganisms (Fig. 17.1; Hallmann et al.<br />

1997b). Von Tiedemann et al. (1983) described a technique for lignified root<br />

tissue. Sterile water is washed through the vascular system at low pressure<br />

to collect the endophytic microorganisms. To prevent damage of the plant<br />

Fig.17.1. Scholander pressure bomb. From left to right: Pressure bomb apparatus, inserting<br />

therootintothepressurecylinder,collectionofplantsapfromtherootwithasterilePasteur<br />

pipette


17 Isolation Procedures for Endophytic Microorganisms 309<br />

tissue, vacuum or pressure should be carefully increased to the maximum<br />

level tolerated by the respective tissue. In all cases, the plant extract is<br />

streaked onto nutrient media to allow recovery of the endophytic microorganisms.<br />

As with the surface sterilisation procedure, sterility checks need<br />

to be included to exclude the possibility that surface colonisers have been<br />

forced from the plant surface through the vascular system by the applied<br />

vacuum or pressure. This can be done by dipping the plant tissue immediately<br />

before extraction into a suspension with an indicator bacterium<br />

and checking for this bacterium after extraction. Additional sterility can<br />

be assured by surface sterilising the cut surface before applying pressure or<br />

vacuum. A sterile working environment can be easily achieved by treatment<br />

with ethanol.<br />

Comparison of the pressure bomb method with maceration of surface<br />

sterilisedrootsresultedinslightlyhigherbacterialnumbersforthelatter<br />

technique (Hallmann et al. 1997b). However, the pressure bomb technique<br />

recovered a higher number of less commonly detected genera, resulting in<br />

higher indices for bacterial richness and diversity. Gram-positive species,<br />

such as Bacillus spp., were more frequently isolated with the maceration<br />

method than with the pressure method, but Gram-negative genera such as<br />

Pseudomonas and Phyllobacterium were recovered at similar levels using<br />

both methods. These results confirm the previously made assumptions that<br />

the pressure bomb method recovers predominantly colonisers of the vascular<br />

tissue, while the maceration method recovers bacteria of the vascular<br />

as well as the cortical root tissue. An advantage of vacuum and pressure<br />

extraction methods is the time saved by avoiding the time-consuming surface<br />

sterilisation procedure. However, limitations of the pressure bomb<br />

technique were encountered for young, fleshy root tissue of cucumber and<br />

tomato, which collapsed under the applied pressure. Those limitations did<br />

not apply to young roots of cotton, soybean and bean (Hallmann et al.<br />

1997b).<br />

17.5<br />

Media<br />

The recipes for most commonly used microbial growth media are listed<br />

for example on the web page of the German National Resource Centre for<br />

Biological Material: http://www2.dsmz.de/media/media.htm.


310 J. Hallmann et al.<br />

17.5.1<br />

Media for Isolating Bacteria<br />

The choice of the growth medium is crucial as it directly affects the number<br />

and type of endophytic microorganisms that can be isolated from the root<br />

tissue. For endophytic bacteria, commonly used media include tryptic soya<br />

agar (TSA), which supports growth of a broad range of bacteria (Gardner<br />

et al. 1982), R2A for bacteria requiring low levels of nutrients (oligotrophs;<br />

Reasoner and Geldreich 1985), nutrient broth-yeast extract medium for<br />

growth of less selective bacteria (Zinniel et al. 2002), King’s B medium<br />

for growth of Pseudomonads (King et al. 1954; Misaghi and Donndelinger<br />

1990) or SC – originally designed for coryneform bacteria, but also useful<br />

for fastidious endophytic bacteria (McInroy and Kloepper 1995). Comparing<br />

the different media, Elvira-Recuenco and van Vuurde (2000) obtained<br />

significantly higher bacterial densities on 5% TSA than on R2A and SC,<br />

while McInroy and Kloepper (1995) found higher bacterial densities on<br />

R2A and SC than on full strength TSA. However, full strength TSA seems to<br />

be less suitable, due to the fact that its high nutrient concentration allows<br />

fast growing bacteria to overgrow slower growing ones. Compared with<br />

SC medium, plate counts from R2A were more accurate due to less colony<br />

overgrowth and smaller colony size (McInroy and Kloepper 1995).<br />

17.5.2<br />

Media for Isolating Fungi<br />

For isolating endophytic fungi, standard media include PDA (potato dextrose<br />

agar; Philipson and Blair 1957), malt extract–peptone–yeast extract<br />

(Schulz et al. 1995) and biomalt agar (Schulz et al. 1995), but minimal media<br />

such as SNA (synthetic nutrient agar; Boyle et al. 2001) may also be used. To<br />

isolate host-specific fungi, host tissue or extracts thereof may be included<br />

in a minimal medium (Arnold and Herre 2003). To increase the diversity<br />

of fungi isolated, it is wise to use a number of different media for each<br />

host plant, varying factors such as pH, temperature of cultivation, and also<br />

aeration.<br />

17.5.3<br />

Supplements<br />

Antibiotics, fungicides or specific nutrients are commonly added to media<br />

to stimulate or suppress growth of certain microbial groups. In spite of<br />

the fact that fungal growth media are usually at slightly acidic pH values,


17 Isolation Procedures for Endophytic Microorganisms 311<br />

antibacterial agents, e.g. oxytetracycline, streptomycin sulfate, penicillin,<br />

and/or novobiocin (Tsao 1970, Schulz et al. 1993), should always be included<br />

in the isolation media. Sublethal doses of fungicides restrict radial<br />

growth of fungal colonies, preventing overgrowth (Bills 1996). For example,<br />

1−2 mg/l Cyclosporin A (active component of Sandimmune; Novartis,<br />

Basel, Switzerland) may be added to the growth medium to suppress fastgrowing<br />

fungi (Dreyfuss and Chapela 1994). Even at low concentrations<br />

in agar (1−10 mg/l), it causes ascomycetous fungi to become unusually<br />

compact, and restricts growth of mucoraceous fungi (Bills 1996). Another<br />

technique is to “weed” out the ubiquitous fungi with a (hot) inoculating<br />

needle (Dreyfuss and Chapela 1994).<br />

17.5.4<br />

Selective Media<br />

Selective culture techniques are used for microorganisms with specialised<br />

physiological capabilities. For example, nitrogen-free enrichment media is<br />

used for the isolation of endophytic diazotrophs (Hartmann et al. 2000),<br />

while media containing chitin, pectin or cellulose as sole nutrient source<br />

are used to select for fungi and bacteria with chitinolytic, pectinolytic or<br />

cellulytic activity, respectively (Chernin et al. 1995; Bills 1996; Berg et al.<br />

2002).<br />

17.6<br />

Cultivation-Independent Methods<br />

Cultivation-dependent methods such as plate counts underestimate microbial<br />

numbers as they do not record viable but non-culturable cells, e.g.<br />

obligate biotrophs, and microorganisms with unknown growth requirements.<br />

Non-culturable microorganisms can be detected using molecular<br />

methods. Their DNA sequences can be compared with those in databanks<br />

to identify the microorganisms. Alternatively, RNA may be used to identify<br />

the active genes. In most cases target sequences are amplified using<br />

the polymerase chain reaction (PCR). The 16S rRNA gene (rDNA) has become<br />

a frequently employed phylogenetic marker with which to describe<br />

bacterial diversity in natural environments (Vossbrinck et al. 1987). Similarly,<br />

18S rDNA is frequently employed for determining fungal diversity<br />

(Kowalchuk et al. 1997; Smalla 2004). However, since it is not very variable,<br />

it may lead to amplification of metazoa (Zuccaro et al. 2003), oomycetes<br />

and diatoms (Nikolcheva et al. 2003). Thus, it is usually more reliable to<br />

amplify 28S rDNA or the internal transcribed spacer (ITS) regions (Schulz<br />

and Boyle 2005).


312 J. Hallmann et al.<br />

Fingerprinting techniques on the basis of PCR-amplified 16S rDNA (bacteria)<br />

or 18S and 28S rDNA (fungi) genes allow for analysis of the structure<br />

of the whole microbial community and their spatial and temporal<br />

variations in relation to environmental factors (Smalla 2004; Zuccaro et<br />

al. 2003; Nikolcheva et al. 2003; Nikolcheva and Bärlocher 2005). Such<br />

techniques include denaturing or temperature gradient gel electrophoresis<br />

(DGGE/TGGE; Kowalchuk et al. 1997; Heuer and Smalla 1997; Zuccaro<br />

2003), terminal restriction fragment length polymorphism (T-RFLP; Liu<br />

et al. 1997; Nikolcheva et al. 2003), and PCR-single-strand-conformationpolymorphism<br />

(SSCP; Schwieger and Tebbe 1998). All these techniques<br />

have been successfully applied to the analysis of endophytic communities,<br />

e.g. T-RFLP (Reiter et al. 2002; Krechel et al. 2002), SSCP (A. Krechel et al.<br />

unpublished data) and DGGE (Garbeva et al. 2001) to characterise endophytes<br />

of potato, DGGE for those from marrum grass (Kowalchuk et al.<br />

1997), citrus plants (Araújo et al. 2002), and fungi associated with decaying<br />

leaves (Nikolcheva and Bärlocher 2005) and algae (Zuccaro et al. 2003).<br />

Using eubacterial primers, it has been estimated that a population must<br />

representabout1%ofthetotalcommunitytobedetectableinafingerprint<br />

(Smalla 2004). Group-specific primers are available for Burkholderia<br />

(Salles et al. 2001), Proteobacteria (Sessitsch et al. 2002), actinomycetes<br />

(Heuer et al. 1997) and many other taxa, allowing a sensitive analysis of the<br />

microbial community. Prominent bands can be excised, cloned and used<br />

for sequence determination in order to obtain further information as to<br />

the identity (Zuccaro et al. 2003) and phylogeny of dominant ribotypes, as<br />

shown for endophytic bacteria of citrus rootstocks by Araújo et al. (2002).<br />

Another cultivation-independent approach is cloning of 16S rDNA, 18S<br />

rDNA or 28S rDNA fragments amplified from community or environmental<br />

DNA, with subsequent sequencing. This approach was applied by Sessitsch<br />

et al. (2002) to the analysis of diversity of potato-associated endophytes.<br />

However, DNA analysis does not generally allow conclusions regarding the<br />

metabolic activity of members of the microbial community or their gene<br />

expression to be drawn. This information may be obtained only through<br />

analysis of RNA (Griffith et al. 2000). Different methodological approaches<br />

have therefore been developed to link information on metabolic activity<br />

to certain ribotypes, such as the incorporation of bromide oxyuridine<br />

(BrdU) or 13 C into the nucleic acids of growing cells (Borneman 1999). Furthermore,<br />

fluorescence in situ hybridization (FISH) as well as marker and<br />

reporter genes are valuable tools with which to study the metabolic activity<br />

of endophytes (Martinez-Inigo et al. 2003; see Chap. 18 by Bloemberg et al.).<br />

However, problematic with molecular results is that: (1) the databanks<br />

are far from complete, and (2) the entries are not always accurate; in the case<br />

of fungi, 20% of entries are estimated to be false (Bridge et al. 2003). Thus, in<br />

order to determine the taxon of a particular microorganism it is important


17 Isolation Procedures for Endophytic Microorganisms 313<br />

not to rely only on one method (e.g. Zuccaro et al. 2003; Nikolcheva and<br />

Bärlocher 2005), but to evaluate multiple characteristics, i.e. morphology<br />

and physiology as well as molecular results.<br />

17.7<br />

Quantification of Colonisation<br />

None of the methods for quantifying the degree of colonisation of endophytic<br />

microorganisms within their hosts is optimal. Colonisation can<br />

be quantified using visual methods, e.g. by direct counts of infections<br />

(Stone 1987); however, this is difficult for bacteria and yeasts. With fungi,<br />

biomass can be correlated with the concentration of fungal specific ergosterol<br />

(Newell et al. 1988; Weete and Ghandi 1996). This method may give<br />

variable results because ergosterol concentration varies with the age of the<br />

mycelium (Olsson et al. 2003). Phospholipid fatty acids have been used for<br />

quantification, but their concentration varies between fungal (Olsson et al.<br />

2003) and bacterial (Olsson and Persson 1999) genera. Fungal biomass can<br />

also be measured using monoclonal antibodies. Here, the difficulty lies in<br />

developing an antibody specific enough to accurately quantify single endophytic<br />

taxa. Real-time PCR (Vaitilingom et al. 1998; Schena et al. 2004) is<br />

presumably the most accurate method for quantifying microbial colonisation<br />

within the host and has been successfully employed, e.g. by Winton et<br />

al. (2002), to quantify the density of fungal colonisation by Phaeocrytopus<br />

gaeumannii within the needles of Douglas-fir and by Oliveira and Vallim<br />

(2002) to quantify colonisation of the bacterial pathogen Xylella fastidiosa<br />

in the leaves of citrus trees.<br />

17.8<br />

Conclusions<br />

The decision as to which isolation procedure is best for a given host depends<br />

on its physical niche, on factors that are plant-dependent, e.g. species, age<br />

and tissue, but also on the available resources and the total number of<br />

samples to be processed (Hallmann et al. 1997a). A polyphasic approach<br />

as suggested in Fig. 17.2 is recommended for the analysis of endophytic<br />

communities. Combinations of different techniques, e.g. surface sterilisation,<br />

vacuum and pressure extraction, cultivation independent (molecular)<br />

methods and cultivation dependent (isolation and morphological) identification,<br />

increase the likelihood of completely analysing microbial diversity,<br />

and consequently also enhancing our understanding of the interactions of<br />

microorganisms with the roots of their plant hosts.


314 J. Hallmann et al.<br />

Fig.17.2. Analysis of endophytic communities using a polyphasic approach<br />

<strong>References</strong><br />

Ahlich K, Sieber T (1996) The profusion of dark septate endophytic fungi in nonectomycorrhizal<br />

fine roots of forest trees and shrubs. New Phytol 132:259–270<br />

Araújo WL, Maccheroni W, Aguilar-Vildoso CI, Barroso PAV, Saridakis HO, Azevedo JL<br />

(2001) Variability and interactions between endophytic bacteria and fungi isolated<br />

from leaf tissue of citrus rootstocks. Can J Microbiol 47:229–236<br />

Araújo WL, Marcon J, Maccheroni W, Van Elsas JD, Van Vuurde JW, Azevedo JL (2002) Diversity<br />

of endophytic bacterial populations and their interaction with Xylella fastidiosa<br />

in citrus plants. Appl Environ Microbiol 68:4906–4914<br />

Arnold AE, Herre EA (2003) Canopy cover and leaf age affect colonisation by tropical fungal<br />

endophytes: ecological pattern and process in Theobroma cacao (Malvaceae). Mycologia<br />

95:388–398<br />

Bell CR, Dickie GA, Harvey WLG, Chan JWYF (1995) Endophytic bacteria in grapevine.<br />

Can J Microbiol 41:46–53<br />

Berg G, Roskot N, Steidle A, Eberl L, Zock A, Smalla K (2002) Plant-dependent genotypic and<br />

phenotypic diversity of antagonistic rhizobacteria isolated from different Verticillium<br />

host plants. Appl Environ Microbiol 68:3328–3338<br />

Bills GF (1996) Isolation and analysis of endophytic fungal communities from woody plants.<br />

In: Redlin SC, Carris LM (eds) Endophytic fungi in grasses and woody plants: systematics,<br />

ecology and evolution. APS Press, St. Paul, MN, pp 31–65


17 Isolation Procedures for Endophytic Microorganisms 315<br />

Borneman J (1999) Culture-independent identification of microorganisms that respond to<br />

specified stimuli. Appl Environ Microbiol 65:3398–3400<br />

Boyle C, Götz M, Dammann-Tugend U, Schulz B (2001) Endophyte–host interactions III.<br />

Local vs. systemic colonisation. Symbiosis 31:259–281<br />

Bridge PD, Roberts PJ, Spooner BM, Panchal G (2003) On the unreliability of published<br />

DNA sequences. New Phytol 160:43–48<br />

Cao LX, You JL, Zhou SN (2002) Endophytic fungi from Musa acuminata leaves and roots<br />

in South China. World J Microbiol Biotechnol 18:169–171<br />

Carroll GC (1995) Forest endophytes: pattern and process. Can J Bot 73(S1):1316–1324<br />

Chernin L, Ismailov Z, Haran S, Chet I (1995) Chitinolytic Enterobacter agglomerans antagonistic<br />

to fungal plant pathogens. Appl Environ Microbiol 61:85–92<br />

Cohen SD (1999) Technique for large scale isolation of Discula umbrinella and other foliar<br />

endophytic fungi from Quercus species. Mycologia 91:917–922<br />

Coombs JT, Franco CMM (2003a) Isolation and identification of actinobacteria from surfacesterilised<br />

wheat roots. Appl Environ Microbiol 69:5603–5608<br />

Coombs JT, Franco CMM (2003b) Visualization of an endophytic Streptomyces species in<br />

wheat seed. Appl Environ Microbiol 69:4260–4262<br />

Crous PW, Petrini O, Marais GF, Pretorius ZA, Rehder F (1995) Occurrence of fungal<br />

endophytes in cultivars of Triticum aestivum in South Africa. Mycoscience 36:739–752<br />

DeWit PJGM, Spikman G (1982) Evidence for the occurrence of race and cultivar-specific<br />

elicitors of necrosis in intercellular fluids of compatible interactions of Cladosporium<br />

fulvum and tomato. Physiol Plant Pathol 21:1–11<br />

Dong Z, Canny MJ, McCully MW, Roboredo MR, Cabadilla CF, Ortega E, Rodés R (1994)<br />

A nitrogen-fixing endophyte of sugarcane stems. Plant Physiol 105:1139–1147<br />

Dreyfuss MM, Chapela IH (1994) Potential of fungi in the discovery of novel, low-molecular<br />

weight pharmaceuticals. In: Gullo VP (ed) The discovery of natural products with<br />

therapeutic potential, Butterworth-Heinemann, Stoneham, MA, pp 49–80<br />

Elvira-Recuenco M, Vuurde van JWL (2000) Natural incidence of endophytic bacteria in<br />

pea cultivars under field conditions. Can J Microbiol 46:1036–1041<br />

Fisher PJ, Petrini O (1992) Fungal saprobes and pathogens as endophytes of rice (Oryza<br />

sativa L.). New Phytol 120:137–143<br />

Gagné S, Richard C, Rousseau H, Antoun H (1987) Xylem-residing bacteria in alfalfa roots.<br />

Can J Microbiol 33:996–1005<br />

Garbeva P, van Overbeek LS, van Vuurde JWL, van Elsas JD (2001) Analysis of endophytic<br />

bacterial communities of potato by plating and denaturing gradient gel electrophoresis<br />

(DGGE) of 16S rDNA based PCR fragments. Microb Ecol 41:369–383<br />

Gardner JM, Feldman AW, Zablotowicz RM (1982) Identity and behavior of xylem-residing<br />

bacteria in rough lemon roots of Florida citrus trees. Appl Environ Microbiol 43:1335–<br />

1342<br />

Griffith RI, Whitely AS, O‘Donell AG, Bailey MJ (2000) Rapid method for coextraction of<br />

DNA and RNA from natural environments for analysis of ribosomal DNA and rRNAbased<br />

microbial community composition. Appl Environ Microbiol 66:5488–5491<br />

Görke C (1998) Mykozönosen von Wurzeln und Stamm von Jungbäumen unterschiedlicher<br />

Bestandsbegründungen. Bibl Mycol 173:1–462<br />

Hallmann J, Sikora RA (1994) Occurrence of plant parasitic nematodes and non-pathogenic<br />

species of Fusarium in tomato plants in Kenya and their role as mutualistic synergists<br />

for biological control of root-knot nematodes. Int J Pest Manage 40:321–325<br />

Hallmann J, Quadt-Hallmann A, Mahaffee WF, Kloepper JW (1997a) Bacterial endophytes<br />

in agricultural crops. Can J Microbiol 43:895–914<br />

Hallmann J, Kloepper JW, Rodríguez-Kábana R (1997b) Application of the Scholander<br />

pressure bomb to studies on endophytic bacteria of plants. Can J Microbiol 43:411–416


316 J. Hallmann et al.<br />

Hallmann J, Quadt-Hallmann A, Rodríguez-Kábana R, Kloepper JW (1998) Interactions<br />

between Meloidogyne incognita and endophytic bacteria in cotton and cucumber. Soil<br />

Biol Biochem 30:925–937<br />

Hambleton S, Currah RS (1997) Fungal endophytes from the roots of alpine and boreal<br />

Ericaceae. Can J Bot 75:1570<br />

Hartmann A, Chatzinotas A, Assmus B, Kirchhof G (2000) Molecular microbial ecology<br />

studies on diazotrophic bacteria associated with non-legumes with special reference to<br />

endophytic diazotrophs. In: Subba Rao NS, Dommergues YR (eds) Microbial interactions<br />

in agriculture and forestry, vol II. Science, Enfield, pp 1–14<br />

Heuer H, Smalla K (1997) Application of denaturing gradient gel electrophoresis (DGGE)<br />

and temperature gradient gel electrophoresis (TGGE) for studying soil microbial communities.<br />

In: van Elsas JD, Wellington EMH, Trevors JT (eds) Modern soil microbiology.<br />

Dekker, New York, pp 353–373<br />

Heuer H, Krsek M, Baker P, Smalla K, Wellington EMH (1997) Analysis of actinomycete communities<br />

by specific amplification of genes encoding 16S rRNA and gel-electrophoretic<br />

separation in denaturing gradients. Appl Environ Microbiol 63:3233–3241<br />

Holdenrieder O, Sieber T (1992) Fungal associations of serially washed healthy nonmycorrhizal<br />

roots of Picea abies. Mycol Res 96:151–156<br />

Hollis JP (1951) Bacteria in healthy potato tissue. Phytopathology 41:350–367<br />

Kattner D, Schönhar S (1990) Untersuchungen über das Vorkommen mikroskopischer Pilze<br />

in Feinwurzeln optisch gesunder Fichten (Picea abies Karst.) auf verschiedenen Standorten.<br />

Mitt Ver Forstl Standortskde Forstpflanzenzücht 35:39–43<br />

King EO, Ward MK, Raney DE (1954) Two simple media for the demonstration of pyocyanin<br />

and fluorescein. J Lab Clin Med 44:301–307<br />

Kowalchuk GA, Gerards S, Woldendorp JW (1997) Detection and characterisation of fungal<br />

infections of Ammophila arenaria (marram grass) roots by denaturing gradient gel<br />

electrophoresis of specifically amplified 18S rDNA. Appl Environ Microbiol 63:3858–<br />

3865<br />

Krechel A, Faupel A, Hallmann J, Ulrich A, Berg G (2002) Potato-associated bacteria and their<br />

antagonistic potential towards plant-pathogenic fungi and the plant-parasitic nematode<br />

Meloidogyne incognita (Kofoid & White) Chitwood. Can J Microbiol 48:772–786<br />

Liu WT, Marsh TL, Cheng H, Forney LJ (1997) Characterisation of microbial diversity by<br />

determining terminal restriction fragment length polymorphisms of genes encoding<br />

16S rRNA. Appl Environ Microbiol 63:4516–4522<br />

Mahaffee WF, Kloepper JW (1997) Temporal changes in the bacterial communities of soil,<br />

rhizosphere, and endorhiza associated with field-grown cucumber (Cucumis sativus L.).<br />

Microb Ecol 34:210–223<br />

Maifeld D (1998) Endophytische Pilze der Fichte (Picea abies (L.) Karst.) – Neue Aspekte<br />

zur biologischen Kontrolle von Heterobasidion annosum (FR.) Bref. PhD dissertation,<br />

Eberhard-Karls-Universität Tübingen, Germany<br />

Martinez-Inigo MJ, Lboe MC, Garbi C, Martin M (2003) Applicability of fluorescence in<br />

situ hybridisation to monitor target bacteria in soil samples. In: Del Re AAM, Capri E,<br />

Padovani L, Trevisan M (eds) Pesticides in air, plant, soil and water system. Proceedings<br />

of the XII Symposium Pesticide Chemistry, Piacenza, Italy, 4–6 June 2003, pp 609–615<br />

McInroy JA, Kloepper JW (1994) Studies on indigenous endophytic bacteria of sweet corn<br />

and cotton. In: O’Gara F, Dowling DN, Boesten B (eds) Molecular ecology of rhizosphere<br />

microorganisms. VCH, Weinheim, Germany, pp 19–28<br />

McInroy JA, Kloepper JW (1995) Survey of indigenous bacterial endophytes from cotton<br />

and sweet corn. Plant Soil 173:337–342<br />

Misaghi IJ, Donndelinger CR (1990) Endophytic bacteria in symptom-free cotton plants.<br />

Phytopathology 80:808–811


17 Isolation Procedures for Endophytic Microorganisms 317<br />

Musson G, McInroy JA, Kloepper JW (1995) Development of delivery systems for introducing<br />

endophytic bacteria into cotton. Biocontrol Sci Technol 5:407–416<br />

Newell SY, Arsuffi TL, Fallon RD (1988) Fundamental procedures for determining ergosterol<br />

content of decaying plant material by liquid chromatography. Appl Environ Microbiol<br />

54: 1876–1879<br />

Nikolcheva LC, Bärlocher F (2005) Seasonal and substrate preferences of fungi colonising<br />

leaves in streams: traditional versus molecular evidence. Environ Microbiol 7:270–80<br />

Nikolcheva LC, Cockshutt AM, Bärlocher F (2003) Determining diversity of freshwater fungi<br />

on decaying leaves: comparison of traditional and molecular approaches. Appl Environ<br />

Microbiol 69:2548–54<br />

Oberholzer-Tschütscher B (1982) Untersuchungen über endophytische Pilze von Erica<br />

carnea L. PhD dissertation, Swiss Federal Institute of Technology, Zürich, Switzerland<br />

O’Dell TE, Trappe JM (1992) Root colonisation of Lupinus latifolius Agardhl and Pinus<br />

contorta Dougl. by Phialocephala fortinii Wang & Wilcox. New Phytol 124:93–100<br />

Oliveira AC, Vallim MA (2002) Quantification of Xylella fastidiosa from citrus trees by<br />

real-time polymerase chain reaction assay. Phytopathology 92:1048–1054<br />

Olsson S, Persson P (1999) The composition of bacterial populations in soil fractions<br />

differing in their degree of adherence to barley roots. Appl Soil Ecol 12:205–215<br />

Olsson PA, Larsson L, Bago B, Wallander H, van Aarle IM (2003) Ergosterol and fatty acids<br />

for biomass estimation of mycorrhizal fungi. New Phytol 159:1–10<br />

Patriquin DG, Döbereiner J (1978) Light microscopy observations of tetrazolium-reducing<br />

bacteria in the endorhizosphere of maize and other grasses in Brazil. Can J Microbiol<br />

24:734–742<br />

Petersen CA, Emanuel ME, Humphreys GB (1981) Pathway of movement of apoplastic<br />

fluorescent dye tracers through the endodermis at the site of secondary root formation<br />

in corn (Zea mays)andbroadbean(Vicia faba). Can J Bot 59:618–625<br />

Petrini O (1984) Endophytic fungi in British Ericaceae. Trans Br Mycol Soc 83:510–512<br />

Petrini O, Fisher PJ, Petrini LE (1992) Fungal endophytes of bracken (Pteridium aquilinum),<br />

with some reflections on their use in biological control. Sydowia 44:282–293<br />

Pleban S, Ingel F, Chet I (1995) Control of Rhizoctonia solani and Sclerotium rolfsii in the<br />

greenhouse using endophytic Bacillus spp. Eur J Plant Pathol 101:665–672<br />

Philipseon MN, Blair ID (1957) Bacteria in clover root tissue. Can J Microbiol 3:125–129<br />

Quadt-Hallmann A, Hallmann J, Kloepper JW (1997) Bacterial endophytes in cotton: location<br />

and interaction with other plant-associated bacteria. Can J Microbiol 43:254–259<br />

Reasoner DJ, Geldreich EE (1985) A new medium for the enumeration and subculture of<br />

bacteria from potable water. Appl Environ Microbiol 49:1–7<br />

Reiter B, Pfeifer U, Schwab H, Sessitsch A (2002) Response of endophytic bacterial communities<br />

in potato plants to infection with Erwinia carotovora subsp. atroseptica.Appl<br />

Environ Microbiol 68:2261–2268<br />

Salles JF, De Souza FA, van Elsas JD (2001) Molecular method to assess the diversity<br />

of Burkholderia species in environmental samples. Appl Environ Microbiol 68:1595–<br />

1603<br />

Sardi P, Sarachhi M, Quaroni S, Petrolini B, Borgonovi GE, Merli S (1992) Isolation of<br />

endophytic Streptomyces strains from surface-sterilised roots. Appl Environ Microbiol<br />

58:2691–2693<br />

Schena L, Nigro F, Ippolito A, Gallitelli D (2004) Real-time quantitative PCR: a new technology<br />

to detect and study phytopathogenic and antagonistic fungi. Eur J Plant Pathol 110:<br />

893–908<br />

Schulz B, Boyle C (2005) The endophytic continuum. Mycol Res 109:661–687<br />

Schulz B, Wanke U, Draeger S, Aust H-J (1993) Endophytes from herbaceous plants and<br />

shrubs: effectiveness of surface sterilisation methods. Mycol Res 97:1447–1450


318 J. Hallmann et al.<br />

Schulz B, Sucker J, Aust HJ, Krohn K, Ludewig K, Jones PG, Döring D (1995) Biologically<br />

active secondary metabolites of endophytic Pezicula species. Mycol Res 99:1007–<br />

1015<br />

Schulz B, Guske S, Dammann U, Boyle C (1998) Endophyte-host interactions II. Defining<br />

symbiosis of the endophyte-host interaction. Symbiosis 25:213–227<br />

Schulz B, Boyle C, Draeger S, Römmert A-K, Krohn K (2002) Endophytic fungi: a source of<br />

biologically active secondary metabolites. Mycol Res 106:996–1004<br />

Schwieger F, Tebbe CC (1998) A new approach to utilize PCR-single-strand-conformation<br />

polymorphism for 16S rRNA gene-based microbial community analysis. Appl Environ<br />

Microbiol 64:4870–4876<br />

Sessitsch A, Reiter B, Pfeifer U, Wilhelm E (2002) Cultivation-independent population<br />

analysis of bacterial endophytes in three potato varieties based on eubacterial and<br />

Actinomycetes-specific PCR of 16S rRNA genes. FEMS Microbiol Ecol 39:23–32<br />

Shishido M, Loeb BM, Chanway CP (1995) External and internal root colonisation of lodgepole<br />

pine seedlings by two growth-promoting Bacillus strains originated from different<br />

root microsites. Can J Microbiol 41:707–713<br />

Sieber TN (2002) Fungal root endophytes. In: Waisel Y, Eshel A, Kafkafi U (eds) Plant roots:<br />

the hidden half. Dekker, New York, pp 887–917<br />

Sieber TN, Riesen TK, Müller E, Fried PM (1988) Endophytic fungi in four wheat cultivars<br />

(Triticum aestivum L.) differing in resistance against Stagonospora nodorum (Berk.)<br />

Berk. J Phytopathol 122:289–306<br />

Skipp RA, Christensen MJ (1989) Fungi invading roots of perennial rye grass (Lolium<br />

perenne L.) in pasture. N J Agric Res 32:423–431<br />

Smalla K (2004) Culture-independent microbiology. In: Bull AT (ed) Microbial diversity and<br />

bioprospecting. ASM, Washington DC, pp 88–99<br />

Sriskandarajah S, Kennedy IR, Yu DG, Tchan YT (1993) Effects of plant growth regulators<br />

on acetylene-reducing associations between Azospirillum brasilense and wheat. Plant<br />

Soil 153:165–178<br />

Stone JK (1987) Initiation and development of latent infections by Rhabdocline parkeri on<br />

Douglas-fir. Can J Bot 65:2614–2621<br />

Stoyke G, Currah RS (1991) Endophytic fungi from the mycorrhizae of alpine ericoid plants.<br />

Can J Bot 69:347–352<br />

Strobel GA (2003) Endophytes as sources of bioactive products. Microbes Infect 5:535–544<br />

Sturz AV (1995) The role of endophytic bacteria during seed piece decay and potato tuberization.<br />

Plant Soil 175:257–263<br />

Tiedemann A von, Wolf G, Wilbert S (1983) Eine einfache Methode zur gezielten Isolierung<br />

von gefäßbesiedelnden Mikroorganismen aus holzigen Pflanzen. Phytopathol Z<br />

107:87–91<br />

Tsao PH (1970) Selective media for isolation of pathogenic fungi. Annu Rev Phytopathol<br />

8:157–186<br />

Vaitilingom M, Gendre F, Brignon P (1998) Direct detection of viable bacteria, molds, and<br />

yeasts by reverse transcriptase PCR in contaminated milk samples after heat treatment.<br />

Appl Environ Microbiol 64:1157–1160<br />

Vossbrinck CG, Maddox JV, Friedman S, Debrunner-Vossbrinck BA, Woese CG (1987)<br />

Ribosomal RNA sequence suggests microsporidia are extremely ancient eukaryotes.<br />

Nature 326:411–414<br />

Werner C, Petrini O, Hesse M (1997) Degradation of the polyamine alkaloid aphelandrine by<br />

endophytic fungi isolated from Aphelandra tetragona. FEMS Microbiol Lett 155:147–153<br />

Weete JD, Ghandi SR (1996) Biochemistry and molecular biology of fungal sterols. In:<br />

Brambl R, Marzluf G (eds) The Mycota III: Biochemistry and molecular biology.<br />

Springer, Berlin Heidelberg New York, pp 421–438


17 Isolation Procedures for Endophytic Microorganisms 319<br />

Winton LM, Stone JK, Watrud LS, Hansen EM (2002) Simultaneous one-tube quantification<br />

of host and pathogen DNA with real-time polymerase chain reaction. Phytopathology<br />

92:112–116<br />

ZinnielDK,LambrechtP,HarrisNB,FengZ,KuczmarskiD,HigleyP,IshimaruCA,Arunakumari<br />

A, Barletta RG, Vidaver AK (2002) Isolation and characterisation of endophytic<br />

colonising bacteria from agronomic crops and prairie plants. Appl Environ Microbiol<br />

68:2198–2208<br />

Zuccaro A, Schulz B, Mitchell J (2003) Molecular detection of ascomycetes associated with<br />

Fucus serratus. Mycol Res 107:1451–1466


18<br />

18.1<br />

Introduction<br />

Microbial Interactions with Plants:<br />

a Hidden World?<br />

Guido V. Bloemberg, Margarita M. Camacho Carvajal<br />

Endophytic microorganisms, e.g. bacteria and fungi, are not only present<br />

within the plant, but may also colonise epiphytically before and during<br />

infection. Some of these organisms have important functions for the plant,<br />

such as nitrogen fixation (e.g. Rhizobiaceae, Herbaspirillum, Azoarcus),<br />

protection against pathogens (e.g. Pseudomonas), and provision of essential<br />

nutrients from the soil (e.g. mycorrhizae), but they can also be parasitic<br />

(e.g. Agrobacterium). Stages in the development of an interaction between<br />

endophyte and plant include attachment to the plant surface, infection<br />

and invasion, colonisation within the plant’s tissues, proliferation (including<br />

sporulation), expression of various phenotypic traits, and spreading<br />

to other plant individuals. Several approaches can be taken to unravel the<br />

complex interactions between the plant and its endophytes. Genetics and<br />

measurement of enzymatic activities are powerful tools with which to discover<br />

the genes and traits involved in these interactions. However, these<br />

processes can only be understood in detail if the microorganisms and the<br />

expression of relevant genes can be visualised on and in the plant, where<br />

the endophytes encounter different tissues and conditions. Visualisation<br />

of the interactions between plant and microorganism on the cellular level<br />

provides a more detailed understanding of the basic principles of how endophytes<br />

function. In this chapter, methods of microscopy and techniques<br />

useful for the analysis of the spatio-temporal behaviour of endophytes<br />

on and in the plant will be evaluated, with emphasis on root-colonising<br />

microorganisms.<br />

Guido V. Bloemberg, Margarita M. Camacho Carvajal: Leiden University, Institute<br />

of Biology, Wassenaarseweg 64, 2333AL Leiden, The Netherlands, E-mail:<br />

bloemberg@rulbim.leidenuniv.nl<br />

Margarita M. Camacho Carvajal: Laboratory for Physiological Chemistry and Centre for<br />

Biomedical Genetics, Utrecht 3584CG, The Netherlands (Present Address)<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


322 G.V. Bloemberg, M.M.C. Carvajal<br />

18.2<br />

Microscopic Techniques for Studying<br />

Plant-Microbe Interactions<br />

Various techniques of microscopy are available for visualising microorganisms<br />

in the plant environment. We will briefly discuss some of the available<br />

techniquesandgiveexamplesofhowthesetechniqueshavebeenvaluable<br />

in studying the interactions of microorganisms with the plant and its<br />

microflora.<br />

18.2.1<br />

Light Microscopy and Enzymatic Reporters<br />

Light microscopes belong to the standard equipment of every microbiology<br />

laboratory and are relatively inexpensive. Several reporter genes with<br />

enzymatic functions such as β-glucuronidase (GUS) encoded by gusA and<br />

β-galactosidase encoded by lacZ are used to visualise bacterial cells and<br />

gene expression (Sambrook and Russel 2001). The advantages of these<br />

reporters are their comparatively low costs and high sensitivity. A clear<br />

disadvantage of using gusA or lacZ as reporters is that plants have to be<br />

fixed, and staining reagents have to penetrate into the plant to reach the<br />

bacteria, which is time consuming, can make detection less efficient, and<br />

is artefact-prone. Another disadvantage can be background staining of the<br />

plant tissue, which should be checked for before using these reporters.<br />

We have used lacZ as a marker to facilitate visualisation of colonisation<br />

by single bacterial cells and bacterial microcolonies of the rhizosphere<br />

of both tomato and Arabidopsis thaliana, and to monitor gene expression<br />

(Fig. 18.1). No background β-galactosidase activity from tomato or<br />

Arabidopsis cells was observed. In particular we focused on the differences<br />

between the rhizosphere colonisation pattern of the biocontrol strain<br />

Pseudomonas fluorescens WCS365 on the model plant Arabidopsis and on<br />

tomato. WCS365 cells preferentially colonised the interjunctions between<br />

root cortex cells and the sites where lateral roots arise (Fig. 18.1A, B). At<br />

such sites, cells are disrupted by the emerging lateral root and nutrients<br />

are released. Cracks in the outer root cortex or plant surface are also very<br />

suitable sites for entering the plant and infecting the internal plant environment.<br />

The colonisation patterns of P. fluorescens WCS365 on tomato<br />

and Arabidopsis are more similar at early (first 3 days after inoculation)<br />

than at later stages, although the number of colony forming units (cfu) per<br />

centimetre of root tip is comparable. It seems that the bacteria around the<br />

root of a 7-day-old tomato plant are embedded in a “mesh” of root hairs<br />

and possibly plant and/or bacterial extracellular material, to which root


18 Microbial Interactions with Plants: a Hidden World? 323<br />

Fig.18.1. A–D LightmicroscopyanalysisoftomatoandArabidopsis thaliana root colonisation<br />

by Pseudomonas fluorescens strain WCS365 constitutively expressing the lacZ reporter<br />

gene encoding β-galactosidase. Sterile seedlings were inoculated with WCS365 2 days after<br />

germination and placed in a gnotobiotic growth system (tomato) or on solid plant nutrient<br />

solution (PNS) agar plates (A. thaliana) for 7 days. After staining for β-galactosidase activity,<br />

roots were examined using light microscopy. A, B Colonisation of the sites of lateral<br />

root emergence from tomato roots. C, D Microcolonies on the root surface of A. thaliana at<br />

the interjunctions of root cells. Bars 100 µm<br />

border cells (BRD) or their excreted proteins (Brigham et al. 1995) may be<br />

major contributors. In contrast, in Arabidopsis the bacteria are located on<br />

the root surface and in between the cell walls of adjacent epidermal cells<br />

(Fig. 18.1C, D). The difference in colonisation patterns between tomato and<br />

Arabidopsis is most likely due to the difference in the rhizosphere environment<br />

created by the two plants. Many plants release BRD from the root tip<br />

into the rhizosphere (Hawes 1990). By definition, BRD are cells that become<br />

dispersed into suspension in response to gentle agitation in water (Hawes<br />

and Lin 1990). The number and properties of BRD vary among different<br />

kinds of plants and are conserved among families (Hawes 1990) Most plants<br />

have BRD but A. thaliana does not (Hawes 1990). BRD are one of the main<br />

components of root exudate (Hawes 1990; Hawes and Lin 1990), which is<br />

the major nutrient source for rhizosphere micro-organisms.<br />

lacZ is not only a useful tool in determining the colonisation pattern of<br />

rhizobacteria in the rhizosphere of different plants and to observe single


324 G.V. Bloemberg, M.M.C. Carvajal<br />

cells within bacterial microcolonies, but can also be used to study bacteriaroot<br />

associations in which the bacteria penetrate deeper into the root tissue,<br />

as was shown for rhizobial cells in infection threads and root nodules (Gage<br />

et al. 1996).<br />

More sophisticated (and more expensive!) light microscopes can be very<br />

valuable for visualising cells without staining, for example endophytic<br />

colonisation and bacteria-fungus interactions. We successfully used phasecontrast<br />

microscopy and differential interference microscopy (DIC) to visualise<br />

the effects of the antifungal phenazine-1-carboxamide on the growth<br />

and hyphal morphology of the plant pathogen Fusarium oxysporum f.sp.<br />

radicis lycopersici (Bolwerk et al. 2003).<br />

18.2.2<br />

Scanning Electron Microscopy<br />

Scanning electron microscopy (SEM) has the advantage of high resolution<br />

and is, therefore, a powerful tool with which to follow the process of seed<br />

and root colonisation by microorganisms. Single bacterial cells and bacterial<br />

microcolonies can be visualised. Using SEM, bacteria were shown<br />

to preferentially colonise grooves on the seed coat and the root surface,<br />

where they formed densely packed microcolonies. Microcolonies are ideal<br />

environments for quorum sensing, which is used by bacteria as a regulatory<br />

process controlling, for example, the production of antifungal metabolites<br />

in the rhizosphere (Fig. 18.2, Chin-A-Woeng et al. 1997). Interestingly,<br />

Fig. 18.2D suggests that the microcolonies are covered by a mucous layer,<br />

possibly consisting of exopolysaccharides, which could form an additional<br />

diffusion barrier.<br />

SEM is also very suitable for showing the morphological differences<br />

within endogenous communities of microorganisms, including obligately<br />

biotrophic bacteria and fungi that cannot be cultured on standard growth<br />

media. A nice example of such a study was given by Fett and Cooke (2003),<br />

who visualised a population of bacteria with different morphologies on<br />

mung bean sprouts.<br />

SEM is ideally suited to visualising the total microflora since it does not<br />

require the tagging of microorganisms with reporters. However, it has the<br />

disadvantage that (1) the living material has to be fixed, and (2) bacterial<br />

species of similar shape and size cannot be differentiated. In addition<br />

the costs of purchasing an electron microscope and its maintenance are<br />

relatively high.


18 Microbial Interactions with Plants: a Hidden World? 325<br />

Fig.18.2. A–D Scanning electron microscopy (SEM) analyses of tomato seed and root colonisation<br />

by Pseudomonas putida WCS358. A Seed coat colonisation 1 day after inoculation.<br />

B Seed coat colonisation monitored 3 days after inoculation. C Attachment on and early<br />

colonisation of the root surface. D Microcolony formed at an interjunction of root cells.<br />

Bars A, B 10 µm; C, D 1 µm. Images provided by T.F.C. Chin-A-Woeng, Leiden University,<br />

Leiden, The Netherlands<br />

18.2.3<br />

Epifluorescence Microscopy and the Application<br />

of Auto-Fluorescent Proteins<br />

Fluorescence microscopy is based on the presence of fluorescent compounds,<br />

including proteins, which, after excitation with light of a certain<br />

wavelength, will emit light of a longer wavelength due to energy loss during<br />

the process of absorption and excitation. Confocal laser scanning microscopy<br />

(CLSM) is a highly sophisticated form of fluorescence microscope.<br />

The use of CLSM for visualising fluorescent molecules results in higher resolution<br />

and lower autofluorescence background compared to traditional<br />

fluorescence microscopy In addition, the resolution and sharpness of the<br />

digital images produced by CLSM can be improved by the use of deconvolution<br />

software that corrects for small defects in the optical lenses. In


326 G.V. Bloemberg, M.M.C. Carvajal<br />

many applications fluorescent tags are coupled to compounds specifically<br />

binding to certain molecules, such as DNA or RNA, in the living cell.<br />

Such compounds are able to diffuse or to be transported into the cell.<br />

Some will diffuse or penetrate differentially through the cell membranes<br />

of Gram positive and Gram negative bacteria. Various fluorescent staining<br />

kits are commercially available. For example, the ViaGram Red kit (Invitrogen<br />

Molecular Probes, Breda, The Netherlands) contains three fluorescent<br />

compounds that offer the possibility to distinguish between Gram positive<br />

and negative bacteria and to test their viability.<br />

Green fluorescent protein (GFP) has become the most frequently used<br />

reporter in the biological sciences since its application as a marker was published<br />

by Chalfie et al. (1994). GFP was isolated from the jellyfish Aequorea<br />

victoria and is extensively applied to mark cells, to visualise proteins within<br />

cells and to monitor gene expression. In contrast to the use of fluorescent<br />

stains,antibodiesorprobestargetedto16SribosomalRNAgenes,theuseof<br />

autofluorescent proteins as markers does not require any preparatory steps<br />

such as fixing with formaldehyde or ethanol, which might affect the biological<br />

material and the biological processes studied. In many cases, fixation<br />

results in death of the living cells. GFP as a reporter has many advantages,<br />

which include (1) stability (due to its barrel protein structure), (2) speciesindependent<br />

application (pro- and eukaryotes), (3) non-invasive analysis<br />

without the need for exogenous substrates or energy, and (4) in vivo monitoring<br />

while preserving the integrity of the living cells. Colour- and optimised<br />

variants of GFP, such as enhanced GFP (EGFP), enhanced cyan fluorescent<br />

protein (ECFP) and enhanced yellow fluorescent protein (EYFP),<br />

with shifted excitation and emission maxima, and increased brightness and<br />

stability have been developed and used for multiple colour imaging (Yang<br />

et al. 1998; Tsien 1998; Matus 1999; Ellenberg 1999). Useful information on<br />

the properties and commercial availability of autofluorescent proteins can<br />

be found on the following website: www.clontech.com/clontech/.<br />

The development of GFP variants with altered excitation and emission<br />

wavelengths and the red fluorescent protein (RFP or DsRed), which was<br />

isolated from the coral Discosoma (Matz et al. 1999), allows for differential<br />

staining of bacterial strains for simultaneous visualisation using CLSM in<br />

one system. We made use of broad host-range cloning vectors for Gram<br />

negative bacteria, which are stably maintained without antibiotic pressure<br />

(Heeb et al. 2000) to tag Pseudomonas and Rhizobium with autofluorescent<br />

proteins encoded by egfp (green), ecfp (cyan), eyfp (yellow), ebfp (blue)<br />

and rfp (red)(Stuurman et al. 2000; Bloemberg et al. 2000). Subsequently,<br />

we were able to visualise simultaneously and clearly differentiate up to<br />

three different bacterial populations at the single cell level in the rhizosphere<br />

(Bloemberg et al. 2000). GFP was shown to be an excellent marker<br />

for monitoring root colonisation at the single cell level by Pseudomonas


18 Microbial Interactions with Plants: a Hidden World? 327<br />

spp. (Bloemberg et al. 1997, 2000), the interactions of Rhizobium with leguminous<br />

plants (Stuurman et al. 2000), the infection process of Fusarium<br />

oxysporum f.sp. radicis lycopersici (Lagopodi et al. 2002) and the interactions<br />

between biocontrol strains of Pseudomonas sp. and F oxysporum f.sp.<br />

radicis lycopersici. (Bolwerk et al. 2003). Besides whole plasmid systems,<br />

valuable transposons have been constructed for integration of gfp into<br />

bacterial chromosomes under constitutive expression (Burlage et al. 1995;<br />

Tombolini et al. 1997; Unge et al. 1997).<br />

18.3<br />

Visualisation of Bacterium-Plant Interactions<br />

Studying microbial communities and their interactions with plants has<br />

been highly facilitated by using combinations of GFP, its colour variants<br />

cyan and yellow fluorescent protein and DsRed as markers. Tomato root<br />

colonisation studies showed that Pseudomonas cells adhere to the root surface<br />

preferentially at the junctions between cells, which are presumed sites<br />

of root exudation, where they proliferate and divide, resulting in the formation<br />

of microcolonies (Bloemberg et al. 1997, 2000; Fig. 18.3A). SEM and<br />

fluorescence microscopy studies revealed that microcolonies are covered<br />

with a mucoid-like layer of unknown origin (Chin-A-Woeng et al. 1997;<br />

Bloemberg et al. 1997). It can be hypothesised that this layer contains bacterial<br />

exopolysaccharides, the production of which is increased in bacterial<br />

biofilms (O’Toole et al. 2000; Sutherland 2001) and which could form a barrier<br />

for signal molecules such as acyl homoserine lactones. Interestingly, we<br />

observed that some plant cells were colonised intracellularly (Fig. 18.3B).<br />

Bacterial cells were also observed close to the openings of the stomata<br />

(Fig. 18.3C), which could form another site of endophytic colonisation, although<br />

we have not observed endophytic colonisation by the Pseudomonas<br />

strains we have used.<br />

Microcolonies are most frequently initiated by one cell, but cells from<br />

external sources can attach to the colony as revealed by a study performed<br />

with a mixture of P. fluorescens WCS365 cells expressing ecfp, egfp and<br />

rfp (Bloemberg et al 2000). It is also expected that cells will detach from<br />

the microcolony to colonise other parts of the growing root. This shows<br />

that bacterial microcolonies are not static, but dynamic entities that can<br />

be invaded by external bacterial cells after the formation of the initial<br />

microcolony. It is of great fundamental interest to find out which genes,<br />

molecules and traits are involved in the attachment and departure of cells.<br />

Studies of an inoculant containing a mixture of P. chlororaphis PCL1391 and<br />

P. fluorescens WCS365 differentially labelled with auto-fluorescent proteins<br />

has shown that (1) predominantly mixed colonies are present on and in


328 G.V. Bloemberg, M.M.C. Carvajal<br />

Fig.18.3. A–F Confocal laser scanning microscopy (CLSM) analysis of tomato surfaces<br />

colonised by Pseudomonas spp. and Fusarium oxysporum f. sp. radicis lycopersici labelled<br />

with autofluorescent proteins. Tomato seedlings were grown upon inoculation in a gnotobiotic<br />

sand system. Plants were taken out after 7 days of growth and examined for the<br />

presence of green fluorescent protein (gfp)- or red fluorescent protein (rfp)-expressing<br />

organisms. A Colonisation of the root surface by P. fluorescens expressing enhanced GFP<br />

(egfp). B Colonisation of the lumen of a root cell. C Colonisation of a stoma. D Simultaneous<br />

imaging of root surface colonisation by a mixture of P. putida PCL1444 expressing egfp<br />

and P. putida PCL1446 expressing rfp. E, F Interactions between P. chlororaphis PCL1391<br />

expressing rfp and the phytopathogenic fungus Fusarium oxysporum f. sp. radicis lycopersici<br />

expressing gfp during colonisation. Bars 10 µm. D courtesy of E. Lagendijk; E, F courtesy of<br />

T. Lagopodi and A. Bolwerk


18 Microbial Interactions with Plants: a Hidden World? 329<br />

older parts of the root, and (2) in a mixture of PCL1391 and WCS365,<br />

PCL1391 predominantly epiphytically colonised the root hairs (Dekkers et<br />

al. 2000), indicating differences in attachment and colonising abilities between<br />

these two closely related Pseudomonas spp. Differential labelling of<br />

strains has also been used to study the colonisation pattern of polyaromatic<br />

hydrocarbon (PAH)-degrading bacteria, which had been isolated for bioremediation<br />

studies. Interestingly, it was shown that these bacterial strains<br />

can form communities on the root as a response to the presence of PAHs<br />

in the soil (Fig. 18.3D).<br />

Results of SEM and CLSM studies show that rhizobacteria, such as<br />

Pseudomonas spp. used for biocontrol, colonise the seed and root surface<br />

at the same positions as phytopathogenic fungi (Chin-A-Woeng et al.<br />

1997; Bloemberg et al. 1997; Tombolini et al. 1999; Lugtenberg et al. 2001;<br />

Lagopodi et al. 2002; Bolwerk et al. 2003; Fig. 18.3). Visualisation is required<br />

to explore and fully understand the interactions between organisms. For<br />

example, visualisation of the relationship among P. fluorescens CHA0, carrot<br />

roots, and mycorrhizal mycelium showed that mucus-producing mutant<br />

strains of CHA0 can better adhere to the root, indicating that acidic extracellular<br />

polysaccharides contribute to root colonisation (Biancotto et al. 2001).<br />

Invasion of plant tissue has been extensively studied for rhizobial infection<br />

that results in the nitrogen-fixing symbiosis with leguminous plants.<br />

Tagging with GFP made it possible to follow early nodulation events, for<br />

example of Sinorhizobium meliloti on alfalfa (Gage et al. 1996), and to follow<br />

rhizobial cells in the infection thread during its growth into the root<br />

cortex. Visualisation of rhizobia in the infection thread made it possible to<br />

calculate the rhizobial growth rate (Gage et al. 1996). GFP tagging was also<br />

successfully applied to study the movement of Rhizobium bacterothe root<br />

nodules (Stuurman et al. 2000). To achieve optimal fluorescence of the bacterial<br />

cells, sectioning of the plant material was required to prevent loss of<br />

light intensity during transmission through the plant tissue. Sectioning of<br />

the plant tissue was also successful in studying the endophytic colonisation<br />

of (1) rice roots and shoots by Herbaspirillum sp. (Elbeltagy et al. 2001) and<br />

(2) Vitis vinifera by the bacterial pathogen Xylella fastidiosa (Newman et<br />

al. 2003). If preservation is preferred or necessary, plant tissue sections can<br />

be fixed with paraformaldehyde, which does not affect the folding and the<br />

fluorescence of GFP (Stuurman et al. 2000; Elbeltagy et al. 2001).<br />

Due to different emission and excitation spectra, a combination of GFP<br />

and DsRed is very suitable for visualisation of two populations of cells<br />

and for providing the possibility of studying competition events. Infection<br />

threads can contain mixed S. meliloti populations that can give rise to mixed<br />

populations of bacteroids in the root nodules (Gage 2002). Since not every<br />

laboratory is equipped with state of the art confocal laser scanning microscopes,<br />

combinations of different reporter genes should be considered.


330 G.V. Bloemberg, M.M.C. Carvajal<br />

Such combinations have been shown to be powerful tools for studying root<br />

colonisation and gene expression in the rhizosphere. Useful constructs<br />

have been made to deliver gfp and gusA in mini-Tn5 transposons or in<br />

plasmids (Ramos et al. 2002) for chromosomal insertion (Xi et al. 1999).<br />

Applying these constructs to the study of Azospirillum brasilense on and in<br />

wheat roots showed that A. brasilense preferentially colonises intercellular<br />

spaces and points of lateral root emergence where it is expected that relatively<br />

large quantities of nutrients will be released from the root cells (Xi<br />

et al. 1999; Ramos et al. 2002). Others have combined immunofluorescence<br />

and an rRNA-targeting probe to monitor the presence of organisms and<br />

metabolic activity in the rhizosphere. For example, P. fluorescens DR54 cells<br />

were analysed in the sugar beet rhizosphere, showing that cells of P. fluorescens<br />

attheroottipweremetabolicallymostactiveandthatbacteria<br />

from the surrounding soil population entered the rhizosphere 2 days after<br />

seed inoculation (Lübeck et al. 2000). Another dual marker system was developed<br />

with gfp and the luxAB genes encoding bacterial luciferase, which<br />

as a biomarker is dependent on cellular energy status. This construct was<br />

used to show that metabolic activity of P. fluorescens SBW25 was detectable<br />

on all parts of wheat and that this strain colonises specific sites of the seed<br />

(Unge et al. 1999; Unge and Jansson 2001).<br />

18.4<br />

Most Recent Developments in Visualising<br />

Plant-Microorganism Interactions<br />

The stability of GFP can be regarded as one of its advantages. However, this<br />

makes GFP unsuitable for studying transient gene expression. GFP derivatives<br />

carrying at their C-terminus amino acid tags for the recognition of<br />

specific proteases have reduced half-life times of GFP to 1–1.5 h (Andersen<br />

et al. 1999). The use of such unstable GFP variants has made it possible<br />

to analyse transient gene expression in the rhizosphere, for example, the<br />

monitoring of ribosomal activity in Pputidacells (Ramos et al. 2000).<br />

These variants have recently been applied to the study of several aspects<br />

of interactions between plants and microorganisms. For example, a system<br />

was constructed for the detection of acyl homoserine lactones (AHL),<br />

showing that quorum sensing and cross talk occur in microcolonies in the<br />

rhizosphere (Andersen et al. 2001; Steidle et al. 2001). Examples of GFPbased<br />

expression systems for studying the interaction of the bacterium<br />

with the plant are given by (1) Leaveau and Lindow (2001), who showed<br />

that foliar growth of Erwinia herbicola on bean is driven by the utilisation<br />

of sugars, e.g. fructose and/or sucrose; and (2) Aldon et al. (2000), who<br />

showed that the strong induction of hrp genes in the presence of the host


18 Microbial Interactions with Plants: a Hidden World? 331<br />

plant cell depends specifically on physicalcontact between the bacterium<br />

and its target cell. The development and application of such reporter systems<br />

contribute significantly to increasing our fundamental knowledge of<br />

bacterial physiology and behaviour on the plant (Leveau and Lindow 2002).<br />

The addition of a constitutively expressed rfpon the reporter construct or<br />

delivered in trans on a second plasmid should make it possible to follow the<br />

presence of bacterial cells and their gene expression. DsRed is specifically<br />

suited to this since it has been isolated from a different organism and<br />

its homology with GFP is extremely low, which excludes the possibility<br />

of recombination when gfp and rfp are present in one cell. DsRed has<br />

a longer folding (maturation) time than GFP, which might reduce or delay<br />

its detection after it is produced. However, an enhanced RFP that matures<br />

rapidly has recently been developed (Sörensen et al. 2003) and will improve<br />

the usefulness of rfp.<br />

In order to identify genes expressed in the plant environment, systems<br />

based on differential fluorescence induction and optical trapping<br />

microscopy have been developed (Allaway et al. 2001). Several genes have<br />

been isolated that have led to new insights into bacterial life in the rhizosphere,<br />

such as the identification of a putative ABC transporter of putrescine<br />

(Allaway et al. 2001). Interestingly, uptake of putrescine from the rhizosphere<br />

environment was identified as an important trait for competitive<br />

root colonisation (Kuiper et al. 2001). New systems for promoter trapping<br />

based on the combination of an antibiotic resistance marker and GFP further<br />

extend the possibilities for researchers to identify genes specifically<br />

expressed in the plant environment (Izallalen et al. 2002).<br />

18.5<br />

Visualisation of Plant-Fungus Interactions<br />

Fungi frequently colonise the internal and external plant environment as<br />

pathogens, mutualists or organisms without apparent effect on the plant.<br />

Fungal structures, including hyphae, fruiting bodies and spores can be visualised<br />

by light- and electron-microscopy with the advantages and disadvantages<br />

of these techniques as discussed above. The use of reporters such<br />

as lacZ or gfp requires genetic transformation, which is much more complicated<br />

for fungi than for bacteria. However, there is a growing number of<br />

fungal species for which transformation methods, including conventional<br />

protoplast transformation, ballistic bombardment and the more recently<br />

developed highly successful Agrobacterium-based transformation method<br />

for filamentous fungi (de Groot et al. 1998), are available Two examples will<br />

be discussed to illustrate the use of GFP in visualisation of fungi interacting<br />

with plants and microorganisms in the plant environment.


332 G.V. Bloemberg, M.M.C. Carvajal<br />

In order to visualise tomato root colonisation and infection processes in<br />

vivo we marked Foxysporumf. sp. radicis-lycopersici, which causes tomato<br />

foot and root rot, with GFP. A protocol to produce Fusarium protoplasts<br />

with the use of a mixture of hydrolytic enzymes was optimised and successfully<br />

applied for cotransformation of the protoplasts with two plasmids,<br />

one of which harboured sgfp, which is an optimised gfp variant (Lagopodi<br />

et al. 2002). This resulted in an efficient tagging and the constitutive sgfp<br />

expression was stable for at least nine subcultures. Homogeneity of the<br />

fluorescent signal was clearly visible in the hyphae as well as in chlamydospores<br />

and conidia. Since, after transformation, the sgfp-harbouring<br />

plasmid is integrated randomly into the chromosome, hyphal morphology,<br />

growth rate and pathogenicity were tested and found to be unaffected in<br />

the transformants tested CLSM was used to analyse colonisation, infection<br />

and disease processes of the isolate in detail on/in tomato roots, including<br />

the following interesting aspects: (1) an overview of the complete colonisation<br />

pattern of the tomato rhizosphere; (2) the very first steps of contact<br />

with the root, which took place in the root hair zone by mingling and attachment<br />

of hyphae to the root hairs, suggesting a chemotactic response<br />

towards the root hairs; (3) the preferential colonisation of the grooves along<br />

the junctions of the epidermal cells, which is similar to the colonisation<br />

patterns of rhizobacteria; (4) the absence of specific infection sites, such as<br />

sites of emergence of secondary roots, root tips or wounded tissue and the<br />

absence of specific infection structures, e.g. appressoria (Lagopodi et al.<br />

2002). This study illustrates the powerful use of GFP as a marker as a noninvasive,<br />

convenient, fast and effective approach for studying plant-fungus<br />

interactions.<br />

Fungi interact with other microorganisms in the plant environment. The<br />

use of different autofluorescent proteins is an excellent tool with which to<br />

distinguish microorganisms from each other and to visualise their interactions.<br />

Visualisation of interactions between phytopathogenic fungi and<br />

biocontrol agents will help to understand these interactions and facilitate<br />

the development of efficient biocontrol applications. We have studied<br />

the interaction between F. oxysporum f. sp. radicis-lycopersici tagged with<br />

GFP and the biocontrol strain P. chlororaphis PCL1391 tagged with DsRed<br />

(Bolwerk et al. 2003). CLSM studies of the single strains showed the following<br />

similarities: (1) the sites of colonisation of the tomato root surface<br />

by Fusarium are strikingly similar to those of the Pseudomonas biocontrol<br />

strains PCL1391 and WCS365; (2) both preferentially colonise sites<br />

between the junctions of two root cells. Studying the interactions between<br />

Pseudomonas biocontrol strains and Fusarium in the rhizosphere by differential<br />

labelling showed that (1) penetration of the fungal hyphae was not<br />

observed where bacterial microcolonies were present, and (2) Pseudomonas<br />

bacteria attached to and colonised Fusarium hyphae (Fig. 18.3F, G). The


18 Microbial Interactions with Plants: a Hidden World? 333<br />

latter could be a novel biocontrol trait and preliminary studies showed that<br />

Pseudomonas strains show a chemotactic response towards Fusarium (De<br />

Weert et al. 2004). (3) Many stress responses in the fungal hyphae were observed<br />

in the presence of the biocontrol agent including increased vacuole<br />

formation, loss of growth directionality, curly growth, “swollen bodies” and<br />

increased hyphal branching. Similar responses could be induced by applying<br />

the purified antifungal metabolite phenazine-1-carboxamide produced<br />

by P. chlororaphis PCL1391.<br />

Studying the molecular basis of the interactions between fungi and bacteria<br />

is an emerging field with great relevance for plant microbiology and<br />

plant pathology. Future studies will benefit greatly from tools developed<br />

for visualisation of these organisms.<br />

18.6<br />

Future Perspectives<br />

The development of highly sophisticated microscopy tools and optimisation<br />

of autofluorescent proteins as reporters are making substantial contributions<br />

to a better fundamental understanding of how microorganisms<br />

interact with plants and the endemic microflora. More reasonably priced<br />

tools for microscopy will facilitate and stimulate such studies in the future.<br />

Visualisation is, however, dependent on whether the microorganims can<br />

be cultured and transformed with genetic constructs that harbour reporter<br />

gene(s). Some important endophytes cannot be cultured outside the plant,<br />

which is severely hampering their study. Discovery and understanding of<br />

plant growth requirements, including chemical signals and physical properties<br />

for providing suitable conditions for culturing, as well as genetic<br />

accessibility of these microorganisms, represents a major challenge.<br />

Acknowledgements. We thank all the members of the Microbiology section<br />

and Gerda Lamers of the Institute of Biology Leiden for their contributions<br />

in valuable discussions for optimising and developing microscopic<br />

techniques, interpretation of results and technical assistance.<br />

<strong>References</strong><br />

Aldon D, Brito B, Boucher C, Genin S (2000) A bacterial sensor of plant cell contact controls<br />

the transcriptional induction of Ralstonia solanacearum pathogenicity genes. EMBO J<br />

19:2304–2314<br />

Allaway D, Schofield NA, Leonard ME, Gilardoni L, Finan TM, Poole PS (2001) Use of differential<br />

fluorescence induction and optical trapping to isolate environmentally induced<br />

genes. Environ Microbiol 3:397–406


334 G.V. Bloemberg, M.M.C. Carvajal<br />

Andersen JB, Sternberg C, Poulsen LK, Bjorn SP, Givskov M, Molin S (1999) New unstable<br />

variants of green fluorescent protein for studies of transient gene expression in bacteria.<br />

Appl Environ Microbiol 64:2240–2246<br />

Andersen JB, Heydorn A, Hentzer M, Eberl L, Geisenberg O, Christensen BB, Molin S,<br />

Givskov M (2001) Gfp-Based N-acyl homoserine-lactone sensor systems for detection<br />

of bacterial communities. Appl Environ Microbiol 67:575–585<br />

Biancotto V, Andreotti S, Balestrini R, Bonfante P, Perotto S (2001) Extracellular polysaccharides<br />

are involved in the attachment of Azospirillum brasilense and Rhizobium leguminosarum<br />

to arbuscular mycorrhizal structures. Eur J Histochem 45:39–49<br />

Bloemberg GV, O’Toole G, Lugtenberg BJJ, Kolter R (1997) Green fluorescent protein as<br />

amarkerforPseudomonas spp. Appl Environ Microbiol 63:4543–4551<br />

Bloemberg GV, Wijfjes AHM, Lamers GEM, Stuurman N, Lugtenberg BJJ (2000) Simultaneous<br />

imaging of Pseudomonas fluorescens WCS365 populations expressing three<br />

different autofluorescent proteins in the rhizosphere; new perspectives for studying<br />

microbial communities. Mol Plant-Microbe Interact 13:1170–1176<br />

Bolwerk A, Lagopodi AL, Wijfjes AHM, Lamers GEM, Chin-A-Woeng TFC, Lugtenberg BJJ,<br />

Bloemberg GV (2003) Interactions in the tomato rhizosphere of two Pseudomonas<br />

biocontrol strains with the phytopathogenic fungus Fusarium oxysporum f. sp. radicislycopersici.<br />

Mol Plant-Microbe Interact 16:983–993<br />

Brigham LA, Woo HH, Nicoll SM, Hawes MC (1995) Differential expression of proteins and<br />

mRNAs from border cells and root tips of pea. Plant Physiol 109:457–463<br />

Burlage RS, Yang ZK, Mehlhorn T (1995) A transposon for green fluorescent protein transcriptional<br />

fusions: applications for bacterial experiments. Gene 173:53–58<br />

Chalfie M, Tu Y, Euskirchen G, Ward WW, Prasher DC (1994) Green fluorescent protein as<br />

a marker for gene expression. Science 263:802–805<br />

Chin-A-Woeng TFC, de Priester W, van der Bij AJ, Lugtenberg BJJ (1997) Description of the<br />

colonisation of a gnotobiotic tomato rhizosphere by Pseudomonas fluorescens biocontrol<br />

strain WCS365, using scanning electron microscopy. Mol Plant-Microbe Interact<br />

10:79–86<br />

De Groot MJ, Bundock P, Hooykaas PJ, Beijersbergen AG (1998) Agrobacterium tumefaciensmediated<br />

transformation of filamentous fungi. Nat Biotechnol 16:839–842<br />

Dekkers LC, Mulders IH, Phoelich CC, Chin-A-Woeng TFC, Wijfjes AHM, Lugtenberg BJJ<br />

(2000) The sss colonisation gene of the tomato-Fusarium oxysporum f.sp. radicis lycopersici<br />

biocontrol strain Pseudomonas fluorescens WCS365 can improve root colonisation<br />

of other wild-type Pseudomonas spp. bacteria. Mol Plant-Microbe Interact 13:1177–1183<br />

De Weert S, Kuiper I, Lagendijk EL, Lamers GEM, Lugtenberg BJJ (2004) Role of chemotaxis<br />

toward fusaric acid in colonisation of hyphae of Fusarium oxysporum f.sp. radicislycopersici<br />

by Pseudomonas fluorescens WCS365. Mol Plant-Microbe Interact 16:1185–<br />

1191<br />

Elbeltagy A, Nishioka K, Sato T, Suzuki H, Ye B, Hamada T, Isawa T, Mitsui H, Minamisawa K<br />

(2001) Endophytic colonisation and in planta nitrogen fixation by a Herbaspirillum sp.<br />

isolated from wild rice species. Appl Environ Microbiol 67:5285–5293<br />

Ellenberg J, Lippincott SJ, Presley JF (1999) Dual-colour imaging with GFP variants. Trends<br />

Cell Biol 9:52–56<br />

Fett WF, Cooke PH (2003) Scanning electron microscopy of native biofilms on mung bean<br />

sprouts. Can J Microbiol 49:45–50<br />

Gage DJ (2002) Analysis of infection thread development using Gfp- and DsRed-expressing<br />

Sinorhizobium meliloti. J Bacteriol 184:7042–7046<br />

Gage DJ, Bobo T, Long SR (1996) Use of green fluorescent protein to visualise the early<br />

events of symbiosis between Rhizobium meliloti and alfalfa (Medicago sativa).JBacteriol<br />

178:7159–7166


18 Microbial Interactions with Plants: a Hidden World? 335<br />

Hawes MC (1990) Living plant cells released from the root cap: a regulator of microbial<br />

populations in the rhizosphere? Plant Soil 129:19–27<br />

Hawes MC, Lin HC (1990) Correlation of pectolytic enzyme activity with the programmed<br />

release of cells from the root cap of Pisum sativum. Plant Physiol 94:1855–1859<br />

Heeb S, Itoh Y, Nishijyo T, Schnider U, Keel C, Wade J, Walsh U, O’Gara F, Haas D (2000)<br />

Small, stable shuttle vectors based on the minimal pVS1 replicon for use in gramnegative,<br />

plant-associated bacteria. Mol Plant-Microbe Interact 13:232–237<br />

Izallalen M, Levesque RC, Perret X, Broughton WJ, Antoun H (2002) Broad-host-range mobilizable<br />

suicide vectors for promoter trapping in gram-negative bacteria. Biotechniques<br />

33:1038–1043<br />

Kuiper I, Bloemberg GV, Noreen S, Thomas-Oates JE, Lugtenberg BJJ (2001) Efficient uptake<br />

of putrescine in the rhizosphere is essential for competitive root colonisation by<br />

Pseudomonas fluorescens strain WCS365. Mol Plant-Microbe Interact 14:1096–1104<br />

Lagopodi AL, Ram AFJ, Lamers GEM, Punt PJ, van den Hondel CAMJ, Lugtenberg BJJ,<br />

Bloemberg GV (2002) Confocal laser scanning microscopical analysis of tomato root<br />

colonisation and infection by Fusarium oxysporum f. sp. radicis-lycopersici using the<br />

green fluorescent protein as a marker. Mol Plant Microb Interact 15:172–179<br />

Leveau JH, Lindow SE (2001) Appetite of an epiphyte: quantitative monitoring of bacterial<br />

sugar consumption in the phyllosphere. Proc Natl Acad Sci USA 98:3446–3453<br />

Leveau JH, Lindow SE (2002) Bioreporters in microbial ecology. Curr Opin Microbiol<br />

5:259–265<br />

Lübeck PS, Hansen M, Sørensen J (2000) Simultaneous detection of the establishment of<br />

seed-inoculated Pseudomonas fluorescens strain Dr54 and native soil bacteria on sugar<br />

beet root surfaces using fluorescense antibody and in situ hybridization techniques.<br />

FEMS Microbiol Ecol 33:11–19<br />

Lugtenberg BJJ, Dekkers L, Bloemberg GV (2001) Molecular determinants of rhizosphere<br />

colonisation by Pseudomonas. Annu Rev Phytopathol 39:461–490<br />

Matus A (1999) GFP in motion CD-ROM. Introduction: GFP illuminates everything. Trends<br />

Cell Biol 9:43<br />

Matz MMV, Fradkov AF, Labas YA, Savitsky AP, Zaraisky AG, Markelov ML, Lukyanov SA<br />

(1999) Fluorescent proteins from non-bioluminescent Anthozoa species. Nat Biotechnol<br />

17:969–973<br />

Newman KL, Almeida RP, Purcell AH, Lindow SE (2003) Use of a green fluorescent strain<br />

for analysis of Xylella fastidiosa colonisation of Vitis vinifera. Appl Environ Microbiol<br />

69:7319–7327<br />

O’Toole G, Kaplan HB, Kolter R (2000) Biofilm formation as microbial development. Annu<br />

Rev Microbiol 54:49–79<br />

Ramos C, Molbak L, Molin S (2000) Bacterial activity in the rhizosphere analyzed at the<br />

single-cell level by monitoring ribosome contents and synthesis rates. Appl Environ<br />

Microbiol 66:801–809<br />

Ramos HJ, Roncato-Maccari LD, Souza FM, Soares-Ramos JR, Hungria M, Pedrosa FO (2002)<br />

Monitoring Azospirillum-wheat interactions using the gfpand gusA genes constitutively<br />

expressed from a new broad-host range vector. J Biotechnol 97:243–52<br />

Sambrook J, Russell DW (2001) Molecular cloning: a laboratory manual. Cold Spring Harbor<br />

Laboratory Press, Cold Spring Harbor, NY<br />

Sörensen M, Lippuner C, Kaiser T, Misslitz A, Aebischer T, Bumann D (2003) Rapidly<br />

maturing red fluorescent protein variants with strongly enhanced brightness in bacteria.<br />

FEBS Lett 552:110–114<br />

Steidle A, Sigl K, Schuhegger R, Ihring A, Schmid M, Gantner S, Stoffels M, Riedel K,<br />

Givskov M, Hartman A, Langebartels C, Eberl L (2001) Visualisation of N-acyl-


336 G.V. Bloemberg, M.M.C. Carvajal<br />

homoserine lactone-mediated cell-cell communication between bacteria colonising the<br />

tomato rhizosphere. Appl Environ Microbiol 67:5761–5770<br />

Stuurman N, Pacios Bras C, Schlaman CHRM, Wijfjes AHM, Bloemberg GV, Spaink HP<br />

(2000) The use of GFP color variants expressed on stable broad-host range vectors to<br />

visualise rhizobia interacting with plants. Mol Plant-Microbe Interact 13:1063–1069<br />

Sutherland IW (2001) The biofilm matrix-an immobilized but dynamic microbial environment.<br />

Trends Microbiol 9:222–227<br />

Tombolini R, Unge A, Davey ME, de Bruijn FJ, Jansson JK (1997) Flow cytometric and<br />

microscopic analysis of GFP-tagged Pseudomonas fluorescens bacteria. FEMS Microbiol<br />

Ecol 22:17–28<br />

Tombolini R, van der Gaag DJ, Gerhardson B, Jansson JK (1999) Colonisation pattern of the<br />

biocontrol strain Pseudomonas chlororaphis MA342onbarleyseedsvisualisedbyusing<br />

green fluorescent protein. Appl Environ Microbiol 65:3674–3680<br />

Tsien RY (1998) The green fluorescent protein. Annu Rev Biochem 67:509–544<br />

Unge A, Jansson JK (2001) Monitoring population size, activity, and distribution of gfpluxAB-tagged<br />

Pseudomonas fluorescens SBW25 during colonisation of wheat. Microb<br />

Ecol 41:290–300<br />

Unge A, Tombolini R, Möller A, Jansson JK (1997) Optimization of GFP as a marker for detection<br />

of bacteria in environmental samples. In: Hastings JW, Kricka LJ, Stanley PE (eds)<br />

Proceedings of the 9th international symposium on bioluminescence and chemiluminescence:<br />

bioluminescence and chemiluminescence: molecular reporting with photons.<br />

Wiley, Chichester, pp 391–394<br />

Unge A, Tombolini R, Molbak L, Jansson JK (1999) Simultaneous monitoring of cell number<br />

and metabolic activity of specific bacterial populations with a dual gfp-luxAB marker<br />

system. Appl Environ Microbiol 65:813–821<br />

Xi C, Lambrecht M, Vanderleyden J, Michiels J (1999) Bi-functional gfp-and gusA-containing<br />

mini-Tn5 transposon derivatives for combined gene expression and bacterial localization<br />

studies. J Microbiol Methods 35:85–92<br />

Yang TT, Sina P, Green G, Kitts PA, Chen YT, Lybarger L, Chervenak R, Patterson GH, Piston<br />

DW, Kain SR (1998) Improved fluorescence and dual color detection with enhanced<br />

blue and green variants of the green fluorescent protein. J Biol Chem 273:8212–8216


19<br />

19.1<br />

Introduction<br />

Application of Molecular Fingerprinting<br />

Techniques to Explore the Diversity<br />

of Bacterial Endophytic Communities<br />

Leo S. van Overbeek, Jim van Vuurde, Jan D. van Elsas<br />

Many bacteria associated with plants are able to penetrate and live as<br />

endophytes in roots and other tissues. Like most bacteria in natural environments,<br />

endophytic bacteria may be non-cultivable, and detectable only<br />

by microscopical and molecular techniques (Garbeva et al. 2001; Araújo<br />

et al. 2002; Sessitsch et al. 2002; Reiter et al. 2003b). Consequently, usage<br />

of molecular techniques reveals higher species diversity than classical<br />

isolation methods. Molecular fingerprinting of bacterial endophytic<br />

populations can expand our knowledge of populations that were previously<br />

inaccessible. So far, little information is available on the possibilities<br />

for employing molecular methods to detect bacterial populations inside<br />

plants. There is, however, overwhelming information from related ecosystems<br />

such as the soil and rhizosphere so that adaptation of methods to<br />

studies on bacterial endophyte communities should be feasible.<br />

In this chapter we will briefly summarise the factors affecting bacterial<br />

endophyte populations in plants and provide an extended account of<br />

the available molecular detection and fingerprinting techniques, including<br />

their integration with other methods, to study such bacterial endophyte<br />

communities. Exploitation of molecular fingerprinting techniques is especially<br />

relevant for studying those endophytic populations that show clear<br />

effects on plant growth and development. Special emphasis will be placed<br />

on agricultural production systems.<br />

Leo S. van Overbeek: Plant Research International B.V, Droevendaalsesteeg 1, 6708 PB,<br />

Wageningen, The Netherlands, E-mail: leo.vanoverbeek@wur.nl<br />

Jim van Vuurde: Plant Research International B.V, Droevendaalsesteeg 1, 6708 PB, Wageningen,<br />

The Netherlands<br />

Jan D. van Elsas: Groningen University, Department of Microbial Ecology, Biological Center,<br />

Kerklaan 30, 9751 NN, Haren, The Netherlands<br />

Soil Biology, Volume 9<br />

Microbial Root Endophytes<br />

B.Schulz,C.Boyle,T.N.Sieber(Eds.)<br />

© Springer-Verlag Berlin Heidelberg 2006


338 L.S.v. Overbeek et al.<br />

19.2<br />

Colonisation by Bacterial Endophytes<br />

Bacterial populations residing on the surface or inside plants are commonly<br />

referred to as plant-associated bacteria. Most of these populations can<br />

colonise both internal and external tissue of plants and therefore a strict<br />

separation between ‘genuine’ endophytes and plant-associated bacteria<br />

cannot always be made. Endophytes thus comprise populations strictly<br />

bound to, and occasionally occupying, internal plant organs.<br />

The initial step in the route to the internal plant system for soil and aerial<br />

bacteria is colonisation of the phytosphere (vis rhizosphere and phyllosphere).<br />

Endophytes occur in the rhizosphere (Sturz 1995; Hallman et al.<br />

1997; Sturz and Nowak 2000) or the phyllosphere (Pillay and Nowak 1997).<br />

Colonisation of internal plant tissue requires adaptation of the bacterial cell<br />

to a highly specific and restrictive environment. Therefore, active colonisation,<br />

systemic spread and reproduction inside plants must be considered<br />

as important features of ‘genuine’ endophytic associations.<br />

Certain plant organs may be environments in which only highly adapted<br />

bacterialspeciesareabletobecomeestablished.Theplantxylemwas<br />

demonstrated to be a selective environment for endophytic bacteria such<br />

as Acetobacter diazotrophicus (Dong et al. 1994), Bacillus pumilus (Benhamou<br />

et al. 1996, 1998), Gluconacetobacter diazotrophicus (Cocking 2003),<br />

Herbaspirillum seropedicae (James et al. 2002), Klebsiella pneumoniae<br />

(Dong et al. 2003) and Serratia marcescens (Gyaneshwar et al. 2001). Sieve<br />

tubes (phloem) are a habitat for a restricted group of pathogens, so called<br />

phytoplasms (Bové and Garnier 2003).<br />

Endophytes can exert diverse effects on the performance of their host<br />

at different stages of growth and under different environmental conditions<br />

(Hallmann et al. 1997; Sturz and Nowak 2000). The effects described<br />

for endophytes inoculated into different plant species may be beneficial<br />

or detrimental [see Chaps. 3 (Kloepper and Ryu) and 4 (Berg and Hallmann)],<br />

although, most often, clear effects cannot be observed. The effect<br />

of bacterial populations associated with plants has been well described for<br />

pathogens and mutualistic symbionts; however, the role of species showing<br />

no obvious effects on plant health during the association may, in general,<br />

be underestimated.<br />

19.3<br />

Shifts of Bacterial Endophyte Communities<br />

Being growing entities, plants have a dynamic interaction with their microbial<br />

inhabitants. Table 19.1 summarises studies on biotic and abiotic factors


19 Molecular fingerprinting of endophytic communities 339<br />

Table 19.1. Factors influencing endophyte colonisation and community structure<br />

Factors affecting<br />

endophytic communitycomposition<br />

in plants<br />

Other microorganisms<br />

Example Reference<br />

Colonisation and nodulation by Rhizobium<br />

leguminosarum in red clover was promoted by<br />

endophytic isolates of Bacillus insolitus, Bacillus brevis<br />

and Agrobacterium rhizogenes<br />

Genotype Higher diversity in root-endophyte composition was<br />

observed in recent versus ancient cultivars of wheat<br />

Differences in endophyte composition and density<br />

were observed in different cotton cultivars<br />

Propagating<br />

material<br />

Agricultural<br />

practices<br />

Localisation<br />

within plant<br />

Enterobacter cloacae applied to sterilised corn seeds<br />

resulted in endophytic colonisation of emerging plants<br />

Optimal colonisation was obtained by inoculation<br />

of in vitro explants with endophytic strain<br />

Pseudomonas sp. PsJN resulting in increased resistance<br />

against Verticillium dahliae in tomato<br />

Gradient in disease-suppressing endophytes observed<br />

from peel to the centre of potato tubers<br />

Crop rotation and tillage management of agricultural<br />

fields resulted in differences in number of disease<br />

suppressive endophytes in potato<br />

Endophytic communities in corn were influenced<br />

by herbicide, feritiliser and compost treatment of soil<br />

Non-nodular nitrogen fixing strain Gluconacetobacter<br />

diazotrophicus was isolated from apoplastic<br />

fluid of sugarcane<br />

Endophytic community structure in potato differed<br />

between the epidermis and internal stem and between<br />

lower and upper stem parts<br />

Temperature Endophytic colonisation of tomato by strain<br />

Pseudomonas sp. PsJN was highest in roots and shoots<br />

at 10 ◦ C and lowest at 30 ◦ C<br />

Sturz and<br />

Christie 1996<br />

Germida and<br />

Siciliano 2001<br />

Adams and<br />

Kloepper 2002<br />

Hinton and<br />

Bacon 1995<br />

Sharma and<br />

Nowak 1998<br />

Sturz et al.<br />

1999<br />

Peters et al.<br />

2003<br />

Seghers et al.<br />

2004<br />

Dong et al.<br />

1994<br />

Garbeva et al.<br />

2001<br />

Pillay and<br />

Nowak 1997<br />

influencing endophytic populations. These studies make it clear that factors<br />

influencing plant growth also influence endophytic populations. The<br />

most important factors governing endophytic community structures are<br />

(1) the environment, (2) characteristics of the host plant species (genotype,<br />

growth stage, tissue type), and (3) endophyte populations already present<br />

in the plant.


340 L.S.v. Overbeek et al.<br />

Inoculation of in vitro plants with endophyte biocontrol strains aimed<br />

to suppress diseases in potato has been proposed by Nowak (1998). The<br />

rational for inoculation of in vitro transplants with endophytes was that<br />

already established populations will have a competitive advantage over<br />

organisms invading from the rhizosphere. The composition of populations<br />

in plants can shift depending on the intricate interplay of different<br />

factors. Eventually, climax populations of endophytes will become established.<br />

19.4<br />

Molecular methods to Study Bacterial Endophytes<br />

Methods to detect beneficial and detrimental endophytic populations are<br />

important in studying the effects of agricultural management on the structure<br />

of endophyte communities and for determining the influence of various<br />

endophyte communities on plant health and crop yield. Genes responsible<br />

for beneficial or detrimental interactions may be the targets for<br />

detection. The structural nitrogen fixation (nif H) gene present in many<br />

nitrogen-fixing species has been detected in bacterial communities inside<br />

and near plants (Reiter et al. 2003a; Tan et al. 2003). Polymerase chain<br />

reaction (PCR)-based detection of genes involved in pathogen suppression<br />

has been developed for pyrrolnitrin and pyoluteorin (De Souza and<br />

Raaijmakers 2003) and ketosynthase (Metsä-Ketelä et al. 2002; Moffitt and<br />

Neilan 2003) genes. The flagellar subunit protein gene, fliC, which codes<br />

for an important protein responsible for host colonisation by the pathogen<br />

Ralstonia solanacearum (Tans-Kersten et al. 2001), was used to establish<br />

a sensitive PCR-based method to detect this pathogen in soil (Schönfeld et<br />

al. 2003). Although most of these detection systems were developed with<br />

the intention of studying bacterial communities in other habitats, they can<br />

easily be applied to study bacterial communities in plants.<br />

The most commonly applied targets for molecular detection in environmental<br />

samples are genes that can be used for taxonomical differentiation,<br />

such as the small and large subunit genes of the ribosomal operon (16S and<br />

23S, respectively), ribosomal RNA (rRNA) genes, and the RNA polymerase<br />

β subunit gene rpoB (Dahllöf et al. 2000). A specific 16S rDNA-based PCR<br />

system aimed at detecting R. solanacearum was, for instance, developed<br />

by Boudazin et al. (1999), whereas a 23S rDNA gene-directed probe was<br />

developed by Wullings et al. (1998), and applied for detection of the same<br />

pathogen using fluorescent in situ hybridisation (FISH) in tomato (Van<br />

Overbeek et al. 2002). A real-time PCR system based on the rpoB gene<br />

has been applied to detect Bacillus anthracis in different environmental<br />

samples (Qi et al. 2001).


19 Molecular fingerprinting of endophytic communities 341<br />

Molecular techniques aimed at detecting specific genes are often restricted<br />

in their application to single targets. Degenerate primers suitable<br />

for PCR detection can overcome this restriction by targeting multiple genes<br />

as, for example, applied to the genes encoding ketosynthase (Metsä-Ketelä<br />

et al. 2002; Moffitt and Neilan 2003) and nif H (Steward et al. 2004). Multiple<br />

sequence detection is crucial for molecular community fingerprinting<br />

of environmental samples, e.g. using PCR coupled with denaturing gradient<br />

gel electrophoresis (PCR-DGGE). Molecular detection techniques are<br />

suitable tools for detecting both cultivable and non-cultivable endophytic<br />

populations in plants.<br />

19.5<br />

Molecular Fingerprinting of Endophyte Communities<br />

19.5.1<br />

Basic Concept of Molecular Fingerprinting<br />

Shifts in endophytic populations and the effect of introduction of selected<br />

endophytes on bacterial communities have recently been studied using<br />

molecular fingerprinting techniques (Garbeva et al. 2001; Sessitsch et al.<br />

2002; Araújo et al. 2002; Reiter et al. 2003b; Conn and Franco 2004; Seghers<br />

et al. 2004). Different methods have been applied, such as PCR-DGGE<br />

(Garbeva et al. 2001; Araújo et al. 2002; Reiter et al. 2003b) and PCR<br />

followed by terminal restriction fragment length polymorphism (PCR-<br />

T-RFLP; Sessitsch et al. 2002). Other molecular fingerprinting techniques,<br />

such as PCR followed by temperature gradient gel electrophoresis (PCR-<br />

TGGE; e.g., Felske et al. 1998b) and PCR-single strand conformational<br />

polymorphism (PCR-SSCP; Schwieger and Tebbe 1998; Schmalenberger<br />

and Tebbe 2003), have not yet been applied to microbial populations in<br />

plants. However, these methods have been successfully applied in studies<br />

of other environments, e.g. bulk and rhizosphere soils.<br />

The principle of all these community fingerprinting techniques, i.e. competitive<br />

PCR amplification of a pool of 16S rRNA gene fragments, is the<br />

same, whereas final analyses of the fragments from environmental samples<br />

differs. The first step in all methods is extraction and purification<br />

of total community nucleic acids. This step is critical, as high molecular<br />

weightDNApureenoughforPCRamplificationisrequired.Secondly,total<br />

microbial community DNA is PCR-amplified using primers spanning fragments<br />

of the 16S rDNA gene. Most commonly, primers that target regions<br />

in the 16S rDNA, which is conserved for all eubacterial species, are used.<br />

However, depending on the focus of the study, primers targeting specific<br />

groups of microorganisms, e.g. eubacteria, fungi or archeabacteria, should


342 L.S.v. Overbeek et al.<br />

be applied. Separation of PCR products is performed in gels with denaturing<br />

gradients (DGGE) or temperature gradients (TGGE), or otherwise,<br />

e.g. SSCP and T-RFLP. A further requirement for PCR-DGGE and PCR-<br />

TGGE is the presence of a GC clamp: a stretch of around 40 nucleotides<br />

of mostly G and C residues. The GC clamp is required for immobilising<br />

denatured DNA fragments within the gel matrix. T-RFLP is recommended<br />

for studying habitats with low species richness (Engebretson and Moyer<br />

2003), which is often the case for internal plant environments (Garbeva et<br />

al. 2001; Sessitsch et al. 2002). T-RFLP has appeared to be more sensitive<br />

for detection of weak bands than PCR-DGGE (Sessitsch et al. 2002; Conn<br />

and Franco 2004).<br />

It is not our intention to describe the details of all methods in full. To<br />

obtain more details about the techniques described in this chapter, we<br />

refer readers interested in this topic to the Molecular microbial ecology<br />

manual (Akkermans et al. 2001). We will focus on critical steps in these<br />

fingerprinting techniques when studying bacterial endophyte populations<br />

as well as the potential offered by these methods. Further, the intricate tasks<br />

and challenges posed by the need to characterise non-cultivable endophytes<br />

will be considered.<br />

19.5.2<br />

Sample Preparation<br />

To discriminate bacteria inside plants (‘genuine’ endophytes) from those<br />

attached or living adjacent to plants, a critical evaluation of methods used<br />

for sample preparation is required. Contamination of samples by bacterial<br />

cells from the outside of plants should be avoided. However, surface sterilisation<br />

of plant parts prior to nucleic acid extraction may not be sufficient<br />

to clean the material of surface nucleic acids. Even minor contamination<br />

may lead to false positive bands in fingerprints due to the sensitivity of<br />

the PCR amplification steps. Moreover, endospores from spore-forming<br />

bacterial species may survive these treatments – these are notorious contaminants<br />

in endophyte studies (Bent and Chanway 2002). Commonly,<br />

surface-sterilised stem parts are incubated on agar medium to check for<br />

the absence of colony growth from the outside (Araújo et al. 2002; Reiter<br />

et al. 2003b; see Chap. 17 by Hallman et al.). Colonies formed from the<br />

plant surface will develop adjacent to the plant part and not within it. This<br />

methodiseffectiveindeterminingtheabsenceofcultivablecellsandspores<br />

originating from the surface that may have survived surface sterilisation.<br />

However, non-cultivable contaminating bacteria will not be detected using<br />

this method.


19 Molecular fingerprinting of endophytic communities 343<br />

For molecular detection, aseptic handling during sample preparation<br />

is extremely important. Surface sterilisation of plant parts followed by<br />

aseptic removal of the outer layer (epidermis) is an effective approach to<br />

remove any traces of external microbial life and thus avoid contamination<br />

of the sample from the outside (Garbeva et al. 2001; Sessitsch et al. 2002).<br />

However, Garbeva et al. (2001) and Sessitsch et al. (2002) concluded that<br />

this approach is feasible only with robust plant parts such as stems and<br />

(potato) tubers but not with fragile structures such as leaves and roots.<br />

A clear distinction between bacteria residing inside and outside these fragile<br />

structures thus cannot easily be made using molecular fingerprinting.<br />

This may be too restrictive when studying ‘genuine’ endophytes. A possible<br />

remedy to selectively isolate DNA and/or RNA from endophytes in finer<br />

structured organs may be a pretreatment with DNase and/or RNase prior<br />

to nucleic acid extraction.<br />

19.5.3<br />

Nucleic Acid Extraction<br />

In principle, both DNA and RNA can be extracted from plant samples<br />

to evaluate endophytic populations. Comparison of fingerprints generated<br />

from DNA and RNA samples may reveal the activity of particular endophyte<br />

populations due to proposed higher ribosomal numbers in metabolically<br />

active cells. Felske et al. (1998b) and Duarte et al. (1998) used this approach<br />

in soil samples. For plant samples, this comparative approach has so far<br />

been applied only once, by Reiter et al. (2003b). The quality and purity<br />

of nucleic acid extracts from plants are the technical constraints for application<br />

of total plant microbial community DNA and RNA in molecular<br />

fingerprinting analyses.<br />

Contamination by polymerase-inhibiting compounds from plants, such<br />

as (poly) phenolics, cannot always be avoided. These vary with the respective<br />

plant species. Community DNA extracts should be diluted in Tris-EDTA<br />

(pH 8) buffer prior to PCR amplification to avoid inhibition of polymerase<br />

activity during PCR.<br />

Although DNA extraction procedures applied in different laboratories<br />

vary, they do not differ fundamentally. Most common procedures include<br />

pulverisation in liquid nitrogen followed by bead beating (Sessitsch et al.<br />

2002; Reiter et al. 2003b) or bead beating of fresh and sliced plant samples<br />

(Garbeva et al. 2001; Araújo et al. 2002). Suspended cells are lysed<br />

in SDS solution and occasionally CTAB (cetyl trimethyl ammonium bromide)<br />

treatment is included to remove plant-derived exopolysaccharides<br />

(Garbeva et al. 2001; Reiter et al. 2003b). DNA is recovered using standard<br />

procedures; i.e. extraction with phenol and chloroform, precipitation


344 L.S.v. Overbeek et al.<br />

in isopropanol and final wash steps in 70% ethanol. Crude extracts may<br />

be further purified using Wizard DNA clean up (Promega, Leiden, The<br />

Netherlands) prior to PCR amplification (Garbeva et al. 2001).<br />

The procedures described above have proven suitable for obtaining nucleic<br />

acids for subsequent molecular fingerprinting. Commercially available<br />

extraction kits, such as the Mo Bio UltraClean soil DNA isolation kit<br />

(Mo Bio Laboratories, BIOzym TC, Landgraaf, The Netherlands), appeared<br />

to be less efficient in the recovery of high quality DNA than both methods<br />

described by Garbeva et al. (2001), and was also our experience in our<br />

laboratories. Although the Mo Bio soil DNA isolation system proved its<br />

validity for recovery of DNA from different soils, it appeared to be limited<br />

to the recovery of plant DNA. Garbeva et al. (2001) also applied an<br />

alternative protocol, in which DNA was extracted from bacterial cells dislodged<br />

by incubating sliced plant parts in buffer. Cells were collected by<br />

centrifugation of the incubation buffer and DNA was extracted from the<br />

cell pellet. The two protocols – DNA extraction from macerated plants and<br />

dislodged cells – were compared by PCR-DGGE analysis (Garbeva et al.<br />

2001). The latter method yielded a clearer pattern with more distinctive<br />

bands in the DGGE gel. A modification of the standard protocol in which<br />

endophytic cells were collected by centrifugation resulted in higher quality<br />

DNA, presumably because there was less contamination with plant-derived<br />

phenolics.<br />

Both methods described by Garbeva et al. (2001) were successfully applied<br />

to DNA extraction from different plants (tomato, leek, chrysanthemum,<br />

lettuce) followed by endophytic fingerprinting with PCR-DGGE in<br />

our laboratories. It is advisable to optimise the nucleic acid extraction protocols<br />

for each newly studied plant species or plant part, although we found<br />

both methods applicable to stems and roots of different plants without further<br />

adaptation. Only optimised sample preparation and DNA extraction<br />

procedures can guarantee successful molecular fingerprinting analysis.<br />

19.5.4<br />

PCR and Molecular Community Fingerprinting<br />

The target sequences present in the total plant-extracted nucleic acid samples<br />

serve as templates for PCR. To assess microbial community structures<br />

in environmental samples, primers that target conserved regions of 16S<br />

rDNA genes are most commonly used. Molecular community fingerprints<br />

will, in principle, reveal all bands from 16S rDNA genes present in total plant<br />

nucleic acid extracts. Each band should represent one taxon. In practice,<br />

bands representing more than one taxon have been reported (Schmalenberger<br />

and Tebbe 2003). Also, single isolates represented by more than one


19 Molecular fingerprinting of endophytic communities 345<br />

band in fingerprints have been observed, as shown for Bacillus sp. strain<br />

Sal1 (Garbeva et al. 2001). The number of individual bands in fingerprints<br />

cannot thus simply be translated to the number of species present in the<br />

plant.<br />

Estimation of the relative population size in molecular fingerprints is<br />

possible by measuring band intensities. However, linear amplification of individual<br />

target sequences in complex DNA extracts will probably not occur.<br />

Therefore, individual bands cannot be used in a straightforward manner<br />

for direct quantification. For additional information about cell numbers of<br />

individual populations, other methods are required. Therefore, molecular<br />

fingerprint analysis is most valuable to demonstrate microbial community<br />

shifts and to compare microbial community structures in different samples.<br />

Plant cell organelles, such as chloroplasts and mitochondria, also possess<br />

16S rDNA genes, and primers targeted to bacteria will also amplify<br />

these genes. Therefore, extra bands may be expected upon fingerprinting<br />

of the amplified products. Identification of individual PCR fragments in<br />

random clone libraries may fail as a result of the high abundance of cell<br />

organelles in total plant DNA extracts. A strategy to exclude 16S rDNA<br />

amplicons of eukaryotic origin is based on pre-amplification with a primer<br />

(799F, Escherichia coli numbering) that does not anneal to chloroplast DNA<br />

(Chelius and Triplett 2001). Clone libraries made in our laboratories by PCR<br />

amplification with primers 799F and 1401R with potato community DNA<br />

extract as template revealed that chloroplast amplicon numbers were lower<br />

in comparison with clone libraries made with eubacterial primers 968F<br />

and 1401R. PCR-DGGE analysis from the same amplicons revealed lower<br />

band intensity at the position where chloroplast bands were expected when<br />

primers 799F and 1401R were applied. Application of primers that target<br />

specific microbial groups also exclude amplicons of chloroplast and mitochondrial<br />

origin (Sessitsch et al. 2002; Reiter et al. 2003b). Group-specific<br />

primers may therefore be best suited for endophyte community structure<br />

analyses.<br />

19.5.5<br />

Group-Specific Molecular Community Fingerprinting<br />

Candidate bacterial endophytes have been characterised for improvement<br />

of plant resistance and increased nutrient acquisition, e.g. by nitrogen fixation,<br />

for several important crops. These cultivable endophytic beneficials,<br />

which will to a great extent control plant health, belong to a limited number<br />

of taxonomic groups. Therefore, molecular tools for detection of these<br />

beneficial populations in plants will certainly gain importance in the near<br />

future. Taxonomic groups representing beneficial endophytes are Aceto-


346 L.S.v. Overbeek et al.<br />

bacter and Gluconacetobacter spp. (Dong et al. 1994; Cocking 2003), Actinomyces<br />

spp. (Zinniel et al. 2002; Castillo et al. 2003; Coombs et al. 2004),<br />

Bacillus spp. (Benhamou et al. 1998; Bacon and Hinton 2002; Reva et al.<br />

2002), Burkholderia spp. (Balandreau et al. 2001), Bradyrhizobium (Chaintreuil<br />

et al. 2000), Enterobacter spp. (Hinton and Bacon 1995), Pseudomonas<br />

spp. (Duijff et al. 1997; Rediers et al. 2003), Rhizobium and Sinorhizobium<br />

spp. (Reiter et al. 2003a) and Serratia spp. [Press et al. 1997; Benhamou<br />

et al. 2000; Tan et al. 2001; Kamensky et al. 2003; see Chaps. 2 (Hallmann<br />

and Berg) and 3 (Kloepper and Ryu)]. However, some of these taxonomic<br />

groups also contain members with non-beneficial or even deleterious properties,<br />

as was the case for certain Serratia marcescens isolates (Gyaneshwar<br />

et al. 2001; Bruton et al. 2003). Endophytic taxa with potentially negative<br />

impacts on human health may belong to the group of Enterobacteriaceae.<br />

Association and endophytic colonisation of plants with human bacterial<br />

pathogens such as Salmonella spp. has recently been reported (Guo et al.<br />

2001, 2002; Cooley et al. 2003; Dong et al. 2003). The lactic acid bacteria, on<br />

the contrary, represent a group of bacteria that are presumed to have positive<br />

effects on human health. A group-specific primer system for lactic acid<br />

bacteria was developed by Heilig et al. (2002). Nevertheless, endophytic<br />

colonisation by lactic acid bacteria has not yet been reported, although it<br />

is known that representatives of this group live in close association with<br />

plants (Ennahar et al. 2003).<br />

Group-specific PCR systems that may be applied for molecular fingerprint<br />

analysis of taxonomically important groups of bacterial endophytes<br />

are presented in Table 19.2. The most important taxons amplified by primers<br />

described in literature are Pseudomonas spp., Bacillus spp, Burkholderia<br />

spp. and Actinomyces spp. Fingerprints from group-specific PCR will give<br />

an impression of all species with close taxonomic relationships to importantknownendophytes.Forpathogensuppression,taxonomicallyrelated<br />

species may be the best competitors for water, nutrients and available<br />

space. For instance, the plant pathogen R. solanacearum was suppressed by<br />

a closely related species, R. picketti, on the rhizoplane of tomato (Shiomi<br />

et al. 1999). Group-specific primers covering all β-proteobacteria, including<br />

Ralstonia sp., appeared to be suitable for studying R. solanacearum<br />

and its taxonomically nearest relatives (Gomes et al. 2001). Primers covering<br />

α-proteobacteria (Gomes et al. 2001) are important because the group<br />

of α-proteobacteria harbours pathogenic species, such as Agrobacterium<br />

tumefaciens, as well as beneficial nitrogen-fixing species such as Rhizobium<br />

and Bradyrhizobium spp.


19 Molecular fingerprinting of endophytic communities 347<br />

Table 19.2. Group-specific primer systems for detection of some endophytic bacterial taxa<br />

(groups)<br />

Bacterial group Nested primer system Reference<br />

Pseudomonas<br />

spp.<br />

Burkholderia<br />

spp.<br />

Actinomyces<br />

spp.<br />

First step: Pseudomonas spp. specific Garbeva<br />

et al. 2004<br />

PsR: 5 ′ -GGTCTGAGAGGATGATCAGT-3 ′<br />

and pSf: 5 ′ -TTAGCTCCACCTCGCGGC-3 ′<br />

Second step: Pseudomonas spp. specific<br />

f968: 5 ′ -AACGCGAAGAACCTTAC- 3 ′ with GC clamp and PsR<br />

First step: eubacterial Reiter et<br />

al. 2003b<br />

8f: 5 ′ -AGAGTTTGATCCTGGCTCCAG-3 ′ and 926r:<br />

5 ′ -CCGTCAATTCCTTT(AG)AGTTT-3 ′<br />

Second step: Pseudomonas spp. specific<br />

8f with CG clamp and PSMGx: 5 ′ -ĆCTTCCTCCCAACTT-3 ′<br />

First step: Burkholderia spp. specific Salles et<br />

al. 2002<br />

Burk3: 5 ′ -CTGCGAAAGCCGGAT-3 ′ and BurkR:<br />

5 ′ -TGCCATACTCTAGCYYGC-3 ′<br />

Second step: Burkholderia spp. specific<br />

Burk3 with GC clamp and r1378:<br />

5 ′ -CGGTGTGTACAAGGCCCGGGAACG-3 ′<br />

First step: Actinomyces spp. specific Heuer et<br />

al. 1997<br />

f243: 5 ′ -GGATGAGCCCGCGGCCTA-3 ′ and r1378<br />

Second step: eubacterial<br />

f984: 5 ′ -AACGCGAAGAACCTTAC-3 ′<br />

with GC clamp and r1378<br />

Bacillus spp. First step: Bacillus spp. specific Garbeva<br />

et al. 2003<br />

Bacf: 5 ′ -GGGAAACCGGGGCTAATACCGGAT-3 ′ and r1378<br />

Second step: eubacterial<br />

f968 with GC clamp and r1378<br />

α-Proteobacteria<br />

β-Proteobacteria<br />

First step: α-proteobacteria specific Gomes et<br />

al. 2001<br />

f203 α:5 ′ -CCGCATACGCCCTACGGGGGAAAGATTTAT-3 ′<br />

and r1494: 5 ′ -CTACGG(T/C)TACCTTGTTACGAC-3 ′<br />

Second step: eubacterial<br />

f984 with GC clamp and r1378<br />

First step: β-proteobacteria specific Gomes et<br />

al. 2001<br />

f203 β:5 ′ -CGCACAAGCGGTGGATGA-3 ′ and r1494<br />

Second step: eubacterial<br />

f984 with GC clamp and r1378


348 L.S.v. Overbeek et al.<br />

19.5.6<br />

Molecular Identification of Species and Genes<br />

Bands in molecular fingerprints reveal neither the identity nor the function<br />

of individual species. For identification of bands of noncultivable species,<br />

individual bands are isolated and, if necessary, the DNA is cloned. Subsequent<br />

sequencing and phylogenetic analysis theoretically enables identification,<br />

provided that the sequences found correspond to known sequences<br />

in the databanks. From the sequence information, new probes can be constructed<br />

for follow-up studies using FISH (e.g. Felske et al. 1998a; Amann<br />

and Ludwig 2000). Noncultivable species may be identified, though their<br />

function and ecological roles will remain unresolved.<br />

Metagenome analysis may represent a new and promising approach in<br />

endophyte research (Rondon et al. 2000). Using this approach the function<br />

of genes from the noncultivable fraction present in different habitats can be<br />

unravelled. Large fragments (>50 kb) are required, prepared from soil or<br />

other environmental DNA samples and cloned into bacterial artificial chromosome<br />

(BAC) vectors. Expression studies and sequence data of genes and<br />

operons of cloned fragments should reveal some of the genetic properties<br />

of noncultivable species in the ecosystem under study. Due to the high incidence<br />

of beneficial bacteria among endophytes and other plant-associated<br />

bacteria commonly observed in many different studies, we expect that the<br />

phytosphere will become an interesting object for metagenome analysis<br />

studies. In the near future, use of the plant-metagenomic approach may<br />

result in new compounds for the development of pharmaceutical and agrochemical<br />

products.<br />

19.6<br />

Integration of Detection Techniques<br />

19.6.1<br />

Polyphasic Approach<br />

Molecular analysis of plant ecosystems has limitations with respect to the<br />

detection, function, activity and ecological behavior of individual populations.<br />

Complementary data are often required, e.g. the use of conventional<br />

techniques of cultivation and microscopy (Van Elsas et al. 1998).<br />

In an unpublished study we used a combination of in situ microbial activity<br />

staining, cultivation and PCR-DGGE to isolate and identify cultivable<br />

and non-cultivable bacteria associated with storage and vascular tissue of<br />

potatotubers.Slicedpotatotuberswereincubatedinwateragarcontaining<br />

triphenyl tetrazolium chloride, a colourless dye that changes into red


19 Molecular fingerprinting of endophytic communities 349<br />

formazan in the presence of microbial activity. Vascular bundles in tubers<br />

appeared to be hot spots for microbial activity and small samples extracted<br />

from the bundles were used for isolation and PCR-DGGE analysis. Two<br />

dominant isolates from the vascular bundle were identified by partial 16S<br />

rDNA gene sequence analysis as Pantoea agglomerans and Bacillus pumilis<br />

Comparison of the bands of these isolates with those of the PCR-DGGE<br />

fingerprints from total vascular community DNA revealed that three dominant<br />

bands did not match any of the isolates. A combination of plating<br />

and PCR-DGGE made it clear that dominant noncultivable species must be<br />

present in the vascular bundles of potato tubers. Additionally, the in situ<br />

metabolic activity staining facilitated sampling of potentially interesting<br />

areas by demonstrating hot spots of microbial activity in tuber tissue.<br />

The application of multiple detection methods will aid in increasing our<br />

knowledge of plant ecosystems. Molecular fingerprinting methods must<br />

be considered as supplementary tools to the already existing panel of detection<br />

techniques. The advantage of molecular fingerprinting techniques<br />

over other techniques is the possibility to detect entire bacterial communities<br />

instead of only cultivable populations. For some ecosystems, such as<br />

soil and water, molecular fingerprinting is nowadays considered a routine<br />

method. A challenge for future endophyte research will be the integration<br />

of molecular fingerprinting methods for cultivable and non-cultivable bacteria<br />

and traditional culture-based endophyte research as complementary<br />

routines.<br />

19.7<br />

Conclusions<br />

Molecular fingerprinting methods are required as tools to supplement conventional<br />

methods in endophyte research to study cultivable and noncultivable<br />

populations in plants. The intricacies of plants make these techniques<br />

difficult to perform. The presence of chloroplast and mitochondrial<br />

DNA in total plant DNA extracts, the abundance of plant DNA versus<br />

bacterial DNA, and the presence of (poly) phenolics may hamper subsequent<br />

analyses. Group (taxon)-specific primer systems can circumvent<br />

co-amplification of 16S rDNA sequences from plant cell organelles. The<br />

plant metagenome concept offers great perspectives for isolating new substances<br />

from plant-associated microorganisms, which may be important<br />

for the pharmaceutical and crop protection industries.


350 L.S.v. Overbeek et al.<br />

<strong>References</strong><br />

Adams PD, Kloepper JW (2002) Effect of host genotype on indigenous bacterial endophytes<br />

of cotton (Gossypium hirsutum L.). Plant Soil 240:181–189<br />

Akkermans ADL, van Elsas JD, de Bruijn JJE (2001) Molecular microbial ecology manual.<br />

Kluwer, Dordrecht<br />

Amann R, Ludwig W (2000) Ribosomal RNA-targeted nucleic acid probes for studies in<br />

microbial ecology. FEMS Microbiol Rev 24:555–565<br />

Araújo WL, Marcon J, Maccheroni W, Van Elsas JD, Van Vuurde JWL, Azevedo JL (2002) Diversity<br />

of endophytic bacterial populations and their interaction with Xylella fastidiosa<br />

in citrus plants. Appl Environ Microbiol 68:4909–4914<br />

Bacon CW, Hinton DM (2002) Endophytic and biological control potential of Bacillus<br />

mojavensis and related species. Biol Control 23:274–284<br />

Balandreau J, Viallard V, Cournoyer B, Coenye T, Laevens S, Vandamme P (2001) Burkholderia<br />

cepacia genomovar III is a common plant-associated bacterium. Appl Environ Microbiol<br />

67:982–985<br />

Benhamou N, Kloepper JW, Quadt-Hallman A, Tuzun S (1996) Induction of defense-related<br />

ultrastructural modifications in pea root tissues inoculated with endophytic bacteria.<br />

Plant Physiol 112:919–929<br />

Benhamou N, Kloepper JW, Tuzun S (1998) Induction of resistance against Fusarium wilt of<br />

tomato by combination of chitosan with an endophytic bacterial strain: ultrastructure<br />

and cytochemistry of the host response. Planta 204:153–168<br />

Benhamou N, Gagné S, Le Quéré D, Dehbi L (2000) Bacterial-mediated induced resistance<br />

in cucumber: beneficial effect of the endophytic bacterium Serratia plymuthica on the<br />

protection against infection by Pythium ultimum. Phytopathology 90:45–56<br />

Bent E, Chanway CP (2002) Potential for misidentification of a spore-forming Paenibacillus<br />

polymyxa isolate as an endophyte by using culture-based methods. Appl Environ<br />

Microbiol 68:4650–4652<br />

Boudazin G, Le Roux AC, Josi K, Labarre P, Jouan B (1999) Design of division specific<br />

primers of Ralstonia solanacearum and application to the identification of European<br />

isolates. Eur J Plant Pathol 105:373–380<br />

Bové JM, Garnier M (2003) Phloem-and xylem-restricted plant pathogenic bacteria. Plant<br />

Sci 164:421–438<br />

Bruton BD, Mitchell F, Fletcher J, Pair SD, Wayadande A, Melcher U, Brady J, Bextine B,<br />

Popham TW (2003) Serratia marcescens, a phloem-colonising, squash bug-transmitted<br />

bacterium: causal agent of cucurbit yellow vine disease. Plant Dis 87:937–944<br />

Castillo U, Harper JK, Strobel GA, Sears J, Alesi K, Ford E, Lin J, Hunter M, Maranta M, Ge H,<br />

Yaver D, Jensen JB, Porter H, Robison R, Millar D, Hess WM, Condron M, Teplow D (2003)<br />

Kakadumycins, novel antibiotics from Streptomyces sp NRRL 30566, an endophyte of<br />

Grevillea pteridifolia. FEMS Microbiol Lett 224:183–190<br />

Chaintreuil C, Giraud E, Prin Y, Lorquin J, Bâ A, Gillis M, Lajudie P, Dreyfus B (2000)<br />

Photosynthetic Bradyrhizobia are natural endophytes of the African wild rice Oryza<br />

breviligulata. Appl Environ Microbiol 66:5437–5447<br />

Chelius MK, Triplett EW (2001) The diversity of archaea and bacteria in association with<br />

the roots of Zea mays L. Microb Ecol 41:252–263<br />

Cocking EC (2003) Endophytic colonisation of plant roots by nitrogen-fixing bacteria. Plant<br />

Soil 252:169–175<br />

Conn VM, Franco CMM (2004) Analysis of the endophytic actinobacterial population<br />

in the roots of wheat (Triticum aestivum L.) by terminal restriction fragment length<br />

polymorphism and sequencing of 16S rRNA clones. Appl Environ Microbiol 70:1787–<br />

1794


19 Molecular fingerprinting of endophytic communities 351<br />

Cooley MB, Miller WG, Mandrell RE (2003) Colonisation of Arabidopsis thaliana with<br />

Salmonella enterica and enterohemorrhagic Escherichia coli O157: H7 and competition<br />

by Enterobacter asburiae. Appl Environ Microbiol 69:4915–4926<br />

Coombs JT, Michelsen PP, Franco CMM (2004) Evaluation of endophytic actinobacteria<br />

as antagonists of Gaeumannomyces graminis var. tritici in wheat. Biol Control<br />

29:359–366<br />

Dahllöf I, Baillie H, Kjelleberg S (2000) rpoB-based microbial community analysis avoids<br />

limitations inherent in 16S rRNA gene intraspecies heterogeneity. Appl Environ Microbiol<br />

66:3376–3380<br />

De Souza JT, Raaijmakers JM (2003) Polymorphisms within the prnDandpltCgenesfrom<br />

pyrrolnitrin and pyoluteorin-producing Pseudomonas and Burkholderia spp. FEMS<br />

Microbiol Ecol 43:21–34<br />

Dong YM, Iniguez AL, Ahmer BMM, Triplett EW (2003) Kinetics and strain specificity of<br />

rhizosphere and endophytic colonisation by enteric bacteria on seedlings of Medicago<br />

sativa and Medicago truncatula. Appl Environ Microbiol 69:1783–1790<br />

Dong ZM, Canny MJ, McCully ME, Roboredo MR, Cabadilla CF, Ortega E, Rodés R (1994)<br />

A nitrogen-fixing endophyte of sugarcane stems – a new role for the apoplast. Plant<br />

Physiol 105:1139–1147<br />

Duarte GF, Rosado AS, Seldin L, Keijzer-Wolters AC, van Elsas JD (1998) Extraction of<br />

ribosomal RNA and genomic DNA from soil for studying the diversity of the indigenous<br />

bacterial community. J Microbiol Methods 32:21–29<br />

Duijff BJ, Gianinazzi-Pearson V, Lemanceau P (1997) Involvement of the outer membrane<br />

lipopolysaccharides in the endophytic colonisation of tomato roots by biocontrol Pseudomonas<br />

fluorescens strain WCS417r. New Phytol 135:325–334<br />

Engebretson JJ, Moyer CL (2003) Fidelity of select restriction endonucleases in determining<br />

microbial diversity by terminal-restriction fragment length polymorphism. Appl<br />

Environ Microbiol 69:4823–4829<br />

Ennahar S, Cai YM, Fujita Y (2003) Phylogenetic diversity of lactic acid bacteria associated<br />

with paddy rice silage as determined by 16S ribosomal DNA analysis. Appl Environ<br />

Microbiol 69:444–451<br />

Felske A, Akkermans ADL, De Vos WM (1998a) In situ detection of an uncultured predominant<br />

Bacillus in Dutch grassland soils. Appl Environ Microbiol 64:4588–4590<br />

Felske A, Wolterink A, Van Lis R, Akkermans ADL (1998b) Phylogeny of the main bacterial<br />

16S rRNA sequences in Drentse A grassland soils (The Netherlands). Appl Environ<br />

Microbiol 64:871–879<br />

Garbeva P, van Overbeek LS, van Vuurde JWL, van Elsas JD (2001) Analysis of endophytic<br />

bacterial communities of potato by plating and denaturing gradient gel electrophoresis<br />

(DGGE) of 16S rDNA based PCR fragments. Microb Ecol 41:369–383<br />

Garbeva P, van Veen JA, van Elsas JD (2003) Predominant Bacillus spp. in agricultural soil<br />

under different management regimes detected via PCR-DGGE. Microb Ecol 45:302–316<br />

Garbeva P, van Veen JA, van Elsas JD (2004) Assessment of the diversity, and antagonism towards<br />

Rhizoctonia solani AG3, of Pseudomonas species in soil from different agricultural<br />

regimes. FEMS Microbiol Ecol 47:51–64<br />

Germida JJ, Siciliano SD (2001) Taxonomic diversity of bacteria associated with the roots<br />

of modern, recent and ancient wheat cultivars. Biol Fert Soils 33:410–415<br />

Gomes NCM, Heuer H, Schönfeld J, Costa R, Mendonca-Hagler L, Smalla K (2001) Bacterial<br />

diversity of the rhizosphere of maize (Zea mays) grown in tropical soil studied by<br />

temperature gradient gel electrophoresis. Plant Soil 232:167–180<br />

Guo X, Chen JR, Brackett RE, Beuchat LR (2001) Survival of Salmonellae on and in tomato<br />

plants from the time of inoculation at flowering and early stages of fruit development<br />

through fruit ripening. Appl Environ Microbiol 67:4760–4764


352 L.S.v. Overbeek et al.<br />

Guo XA, van Iersel MW, Chen JR, Brackett RE, Beuchat LR (2002) Evidence of association of<br />

Salmonellae with tomato plants grown hydroponically in inoculated nutrient solution.<br />

Appl Environ Microbiol 68:3639–3643<br />

Gyaneshwar P, James EK, Mathan N, Reddy PM, Reinhold-Hurek B, Ladha JK (2001) Endophytic<br />

colonisation of rice by a diazotrophic strain of Serratia marcescens. JBacteriol<br />

183:2634–2645<br />

Hallmann J, Quadt-Hallmann A, Mahaffee WF, Kloepper JW (1997) Bacterial endophytes<br />

in agricultural crops. Can J Microbiol 43:895–914<br />

Heilig HGHJ, Zoetendal EG, Vaughan EE, Marteau P, Akkermans ADL, de Vos WM (2002)<br />

Molecular diversity of Lactobacillus spp. and other lactic acid bacteria in the human<br />

intestine as determined by specific amplification of 16S ribosomal DNA. Appl Environ<br />

Microbiol 68:114–123<br />

Heuer H, Krsek M, Baker P, Smalla K, Wellington EMH (1997) Analysis of actinomycete<br />

communities by specific amplification of genes encoding 16S rRNA and gelelectrophoretic<br />

separation in denaturing gradients. Appl Environ Microbiol 63:3233–<br />

3241<br />

Hinton DM, Bacon CW (1995) Enterobacter cloacae is an endophytic symbiont of Corn.<br />

Mycopathologia 129:117–125<br />

James EK, Gyaneshwar P, Mathan N, Barraquio WL, Reddy PM, Iannetta PPM, Olivares FL,<br />

Ladha JK (2002) Infection and colonisation of rice seedlings by the plant growthpromoting<br />

bacterium Herbaspirillum seropedicae Z67. Mol Plant-Microbe Interact<br />

15:894–906<br />

Kamensky M, Ovadis M, Chet I, Chernin L (2003) Soil-borne strain IC14 of Serratia plymuthica<br />

with multiple mechanisms of antifungal activity provides biocontrol of Botrytis<br />

cinerea and Sclerotinia sclerotiorum diseases. Soil Biol Biochem 35:323–331<br />

Metsä-Ketelä M, Halo L, Munukka E, Hakala J, Mäntsälä P, Ylihonko K (2002) Molecular<br />

evolution of aromatic polyketides and comparative sequence analysis of polyketide<br />

ketosynthase and 16S ribosomal DNA genes from various Streptomyces species. Appl<br />

Environ Microbiol 68:4472–4479<br />

Moffitt MC, Neilan BA (2003) Evolutionary affiliations within the superfamily of ketosynthases<br />

reflect complex pathway associations. J Mol Evol 56:446–457<br />

Nowak J (1998) Benefits of in vitro “biotisation” of plant tissue cultures with microbial<br />

inoculants. In Vitro Cell Dev Biol-Plant 34:122–130<br />

Peters RD, Sturz AV, Carter MR, Sanderson JB (2003) Developing disease-suppressive soils<br />

through crop rotation and tillage management practices. Soil Tillage Res 72:181–192<br />

Pillay VK, Nowak J (1997) Inoculum density, temperature, and genotype effects on in vitro<br />

growth promotion and epiphytic and endophytic colonisation of tomato (Lycopersicon<br />

esculentum L) seedlings inoculated with a pseudomonad bacterium. Can J Microbiol<br />

43:354–361<br />

Press CM, Wilson M, Tuzun S, Kloepper JW (1997) Salicylic acid produced by Serratia<br />

marcescens 90-166 is not the primary determinant of induced systemic resistance in<br />

cucumber or tobacco. Mol Plant-Microbe Interact 10:761–768<br />

Qi YA, Patra G, Liang X, Williams LE, Rose S, Redkar RJ, DelVecchio VG (2001) Utilisation of<br />

the rpoB gene as a specific chromosomal marker for real-time PCR detection of Bacillus<br />

anthracis. Appl Environ Microbiol 67:3720–3727<br />

Rediers H, Bonnecarrère V, Rainey PB, Hamonts K, Vanderleyden J, De Mot R (2003)<br />

Development and application of a dapB-based in vivo expression technology system to<br />

study colonisation of rice by the endophytic nitrogen-fixing bacterium Pseudomonas<br />

stutzeri A15. Appl Environ Microbiol 69:6864–6874<br />

Reiter B, Bürgmann H, Burg K, Sessitsch A (2003a) Endophytic nif H gene diversity in<br />

African sweet potato. Can J Microbiol 49:549–555


19 Molecular fingerprinting of endophytic communities 353<br />

Reiter B, Wermbter N, Gyamfi S, Schwab H, Sessitsch A (2003b) Endophytic Pseudomonas<br />

spp. populations of pathogen-infected potato plants analysed by 16S rDNA- and 16S<br />

rRNA-based denaturating gradient gel electrophoresis. Plant Soil 257:397–405<br />

Reva ON, Smirnov VV, Pettersson B, Priest FG (2002) Bacillus endophyticus sp nov., isolated<br />

from the inner tissues of cotton plants (Gossypium sp.).IntJSystEvolMicrobiol<br />

52:101–107<br />

Rondon MR, August PR, Bettermann AD, Brady SF, Grossman TH, Liles MR, Loiacono KA,<br />

Lynch BA, MacNeil IA, Minor C, Tiong CL, Gilman M, Osburne MS, Clardy J, Handelsman<br />

J, Goodman RM (2000) Cloning the soil metagenome: a strategy for accessing the<br />

genetic and functional diversity of uncultured microorganisms. Appl Environ Microbiol<br />

66:2541–2547<br />

Salles JF, De Souza FA, van Elsas JD (2002) Molecular method to assess the diversity of<br />

Burkholderia species in environmental samples. Appl Environ Microbiol 68:1595–1603<br />

Schmalenberger A, Tebbe CC (2003) Bacterial diversity in maize rhizospheres: conclusions<br />

on the use of genetic profiles based on PCR-amplified partial small subunit rRNA genes<br />

in ecological studies. Mol Ecol 12:251–261<br />

Schönfeld J, Heuer H, van Elsas JD, Smalla K (2003) Specific and sensitive detection of<br />

Ralstonia solanacearum in soil on the basis of PCR amplification of fliC fragments. Appl<br />

Environ Microbiol 69:7248–7256<br />

Schwieger F, Tebbe CC (1998) A new approach to utilise PCR-single-strand-conformation<br />

polymorphism for 16s rRNA gene-based microbial community analysis. Appl Environ<br />

Microbiol 64:4870–4876<br />

Seghers D, Wittebolle L, Top EM, Verstraete W, Siciliano SD (2004) Impact of agricultural<br />

practices on the Zea mays L. endophytic community. Appl Environ Microbiol 70:1475–<br />

1482<br />

Sessitsch A, Reiter B, Pfeifer U, Wilhelm E (2002) Cultivation-independent population<br />

analysis of bacterial endophytes in three potato varieties based on eubacterial<br />

and Actinomycetes-specific PCR of 16S rRNA genes. FEMS Microbiol Ecol<br />

39:23–32<br />

Sharma VK, Nowak J (1998) Enhancement of Verticillium wilt resistance in tomato transplants<br />

by in vitro co-culture of seedlings with a plant growth promoting rhizobacterium<br />

(Pseudomonas sp. Strain PsJN). Can J Microbiol 44:806–806<br />

Shiomi Y, Nishiyama M, Onisuka T, Marumoto T (1999) Comparison of bacterial community<br />

structures in the rhizoplane of tomato plants grown in soils suppressive and conducive<br />

towards bacterial wilt. Appl Environ Microbiol 65:3996–4001<br />

Steward GF, Jenkins BD, Ward BB, Zehr JP (2004) Development and testing of a DNA<br />

macroarray to assess nitrogenase (nif H) gene diversity. Appl Environ Microbiol<br />

70:1455–1465<br />

Sturz AV (1995) The role of endophytic bacteria during seed piece decay and potato tuberisation<br />

Plant Soil 179:303–303<br />

Sturz AV, Christie BR (1996) Endophytic bacteria of red clover as agents of allelopathic<br />

clover-maize syndromes. Soil Biol Biochem 28:583–588<br />

Sturz AV, Nowak J (2000) Endophytic communities of rhizobacteria and the strategies<br />

required to create yield enhancing associations with crops. Appl Soil Ecol 15:183–190<br />

Sturz AV, Christie BR, Matheson BG, Arsenault WJ, Buchanan NA (1999) Endophytic bacterial<br />

communities in the periderm of potato tubers and their potential to improve<br />

resistance to soil-borne plant pathogens. Plant Pathol 48:360–369<br />

Tan ZY, Hurek T, Gyaneshwar P, Ladha JK, Reinhold-Hurek B (2001) Novel endophytes of<br />

rice form a taxonomically distinct subgroup of Serratia marcescens.SystApplMicrobiol<br />

24:245–251


354 L.S.v. Overbeek et al.<br />

Tan XY, Hurek T, Reinhold-Hurek B (2003) Effect of N-fertilisation, plant genotype and<br />

environmental conditions on nif H gene pools in roots of rice. Environ Microbiol 5:1009–<br />

1015<br />

Tans-Kersten J, Huang HY, Allen C (2001) Ralstonia solanacearum needs motility for invasive<br />

virulence on tomato. J Bacteriol 183:3597–3605<br />

Van Elsas JD, Duarte GF, Rosado AS, Smalla K (1998) Microbiological and molecular biological<br />

methods for monitoring microbial inoculants and their effects in the soil<br />

environment. J Microbiol Methods 32:133–154<br />

Van Overbeek LS, Cassidy M, Kozdroj J, Trevors JT, van Elsas JD (2002) A polyphasic<br />

approach for studying the interaction between Ralstonia solanacearum and potential<br />

control agents in the tomato phytosphere. J Microbiol Methods 48:69–86<br />

Wullings BA, van Beuningen AR, Janse JD, Akkermans ADL (1998) Detection of Ralstonia<br />

solanacearum, which causes brown rot of potato, by fluorescent in situ hybridisation<br />

with 23S rRNA-targeted probes. Appl Environ Microbiol 64:4546–4554<br />

ZinnielDK,LambrechtP,HarrisNB,FengZ,KuczmarskiD,HigleyP,IshimaruCA,Arunakumari<br />

A, Barletta RG, Vidaver AK (2002) Isolation and characterisation of endophytic<br />

colonising bacteria from agronomic crops and prairie plants. Appl Environ Microbiol<br />

68:2198–2208


Subject Index<br />

Abies alba 304<br />

Abiotic stress 271<br />

Acephala applanata 121<br />

Acer spicatum 182–185<br />

Acetobacter 7<br />

– diazotrophicus 26, 338<br />

Achlya klebsiana 55<br />

Acremonium 163<br />

– alternatum 270, 274<br />

–cf.curvulum 232<br />

– cucurbitacearum 216<br />

– killense 159<br />

– pteridii 166<br />

– strictum 164, 237<br />

Acrocalymma medicaginis 210<br />

Actinoplanes missouriensis 59, 61<br />

Acuminatopyrone 146<br />

Adaptations 4<br />

Adhesive 194<br />

Aequorea victoria 326<br />

Agricultural practices 25<br />

Agrobacterium 321<br />

– tumefaciens 23, 56, 82, 346<br />

Agrobacterium-based<br />

transformation 331<br />

Agrochemicals 117<br />

Air pollutants 118<br />

Allozyme 122<br />

Alnus glutinosa 110, 181, 183–186,<br />

188, 189<br />

α-proteobacteria 346<br />

Altechromones 268, 274<br />

Alternaria alternata 162, 165<br />

Altitude 109<br />

AMF 109, 117, 119<br />

Amycolatopsis meditarranei 56<br />

Anamorph 187<br />

Angiopteris evecta 182,<br />

184–186<br />

Anguillospora<br />

– filiformis 183<br />

– longissima 181, 183<br />

Angular leaf spot 35<br />

Antagonism 54, 56–59, 67, 68<br />

– balanced 172<br />

Antagonistic 53–58, 61, 64–67<br />

Antagonistic bacteria 58<br />

Antagonistic endophytic bacteria used to<br />

control fungal pathogens. BCA<br />

Biocontrol agent 62<br />

Anthostomella aracearum 166<br />

Antibiosis 53, 58, 59, 67<br />

Antibiotic 180, 186<br />

Antifungal 57–60, 66<br />

Antipteris evecta 186<br />

Aphelandra tetragona 303<br />

Apoplastic fluid 307<br />

Appressorium 192, 194, 197<br />

Arabidopsis 43<br />

– thaliana 96, 322<br />

Arbuscular mycorrhiza 119, 209, 248, 266<br />

Arbutoid mycorrhiza 209<br />

Armillaria jezoensis 161<br />

– mellea 160, 161, 168<br />

Armillariella mellea 161<br />

Arthrinium sp. 163, 165<br />

Arthrobacter 155<br />

– ilicis 56<br />

Arthrobotrys 192<br />

– dactyloides 196, 198–200<br />

– ferox 192<br />

– oligospora 193, 197, 198<br />

Articulospora<br />

– antipodea 183<br />

– atra 183<br />

– proliferata 183<br />

– tetracladia 183<br />

Artificial inoculation 16


356 Subject Index<br />

Ascochyta sp. 163, 166<br />

Ascomycetes 193, 208<br />

Ascomycota 219<br />

Aspergillus 163<br />

Assemblages 4<br />

Astroloma pinifolium 212<br />

Asymptomatic 6<br />

Aureobasidium caulivorum 166<br />

Auricularia polytricha 161<br />

Aurofusarin 146<br />

Avicennia<br />

– officinalis 183, 185<br />

Azoarcus 2, 7, 11, 96<br />

Azospirillum 39<br />

– brasilense 330<br />

B trichothecene 144<br />

BAC 348<br />

Bacillispora<br />

– inflata 183<br />

Bacillus 56, 61, 66, 68, 91, 155, 321<br />

– amyloliquefaciens 36<br />

– aquamarinus 56<br />

– cereus 26, 56, 59, 68<br />

– licheniformis 65<br />

– megaterium 22, 56<br />

– polymyxa 94<br />

– pumilus 34, 36, 65, 338, 349<br />

– subtilis 36, 58, 59, 65<br />

Bacteria 154<br />

– nitrogen-fixing 155<br />

Bacterial 338<br />

Bacterial artificial chromosome 348<br />

Bacterial diversity 21<br />

Bacterial endophytes 90<br />

Bacterial identification 16<br />

Bacterial speck 39<br />

Bacterial spectrum 16<br />

Bacteroids 71<br />

Balance of antagonisms 264<br />

Balanced antagonism 7, 219, 269<br />

Balansia 134<br />

Basidiomycetes 192<br />

BCA 355<br />

Bead beating 343<br />

Beauvericins 264<br />

Behaviour 107<br />

– antagonistic 107, 119<br />

– mutualistic 107<br />

– neutral 107, 110, 114<br />

– pathogenic 107, 120<br />

Beneficial 340<br />

Benomyl 118<br />

Benzoxazinoids 145<br />

β-glucoronidase 136<br />

β-proteobacteria 346<br />

Betula<br />

– papyrifera 117, 182–185<br />

– pendula 112, 304<br />

– platyphylla 118<br />

– pubescens 212<br />

Bidirectional carbon transport 269<br />

Biocides 117<br />

Biodiversity 107, 157<br />

Biofilms 327<br />

Biological control 6, 33, 53, 54, 57–61,<br />

64–67, 69, 191<br />

Biomass 115<br />

Biotrophic 147<br />

BioYield 49<br />

Bishop pine 118<br />

Blue mold 37<br />

Bog 229, 232<br />

Bordetella pertussis toxin exporter 82<br />

Bracken 115<br />

Bradyrhizobium japonicum 77<br />

Brassica chinensis var.<br />

parachinensis 38<br />

Bremisia argentifolii 37<br />

Burkholderia 54<br />

– cepacia 26, 65<br />

– solanacearum 61<br />

Butenolide 146<br />

Cadophora finlandia 211, 216<br />

Cajanus cajan 79<br />

Calanthe 155<br />

Callose 197<br />

Calluna vulgaris 236<br />

Calonectria kyotensis 162<br />

Campylospora<br />

– chaetocladia 184<br />

– purvula 181<br />

Capronia 212, 249, 253<br />

Capsicum annuum 38<br />

Capsular polysaccharides 73


Subject Index 357<br />

Carbohydrates 114<br />

Carbon sink 272<br />

Carbon transport 266<br />

Catenaria anguillulae 193<br />

Cell organelles 345<br />

Cell wall appositions 197, 270, 271<br />

Centrifugation 307<br />

Centrospora acerina 181<br />

Cephalanthera austinae 167<br />

Ceratobasidium 156, 166<br />

Ceratocystis fagacearum 91<br />

Ceratorhiza, Epulorhiza 166<br />

Cereal 116, 120<br />

Chaetomium aureum 165<br />

– funicola 164, 265, 274<br />

– homopilatum 162<br />

– subspirale 163, 164<br />

Chaetosphaeria endophytica,<br />

Colletotrichum gloeosporioides 166<br />

Chaetosticta cf. perforata 165<br />

Chaetothyriomycetidae 210, 212<br />

Chemical messengers 117<br />

Chinese cabbage 112<br />

Chitinase 194<br />

Chlamydospores 118, 125, 197<br />

Chlamydosporol 146<br />

Chloridium paucisporum 120, 211, 262<br />

– virescens 165<br />

Chloroplasts 345<br />

Christela<br />

– dentata 182, 184–186<br />

Chrysogine 146<br />

Chytridiomycetes 193<br />

Citrus jambhiri 302<br />

Cladorrhinum foecumdissimum 274<br />

Cladosporium cladosporioides 164<br />

Clavariopsis<br />

– aquatica 184<br />

– azlanii 184<br />

Clavibacter michigenensis 56<br />

Climate 108, 109<br />

Climate change 108<br />

Clone libraries 345<br />

Clover 111, 117<br />

CO2 109<br />

Codinaea parva 162<br />

Coelomycete 210<br />

Coffea arabica 182, 184, 185, 186<br />

Colletotrichum 163–165, 168<br />

– acutatum 165<br />

– crassipes 162–165<br />

– gloeosporioides 38, 163, 166<br />

– magna 6<br />

– orbiculare 36<br />

Colonisation 262, 263, 273<br />

Commensalism 6<br />

Common reed 115<br />

Communities 107, 108, 110, 113–115,<br />

117, 118, 127, 329, 337<br />

Community composition 108<br />

Competition 59, 115, 120<br />

– interspecific 115, 125<br />

Competitive advantage 107<br />

Competivity 109<br />

Confocal laser scanning microscopy 325<br />

Conidia 197<br />

Conidial traps 196<br />

Coniothyrium sp. 166<br />

Continuum 6<br />

Corallorhiza 167<br />

Corn (Zea mays) 90<br />

Cortinarius 286<br />

Cotton 40<br />

Cronartium quercuum f. sp. fusiforme 38<br />

Crop protection 118<br />

Crotalaria juncea 78<br />

Cryptocline 163, 164, 166<br />

Cryptosporiopsis 4, 163–166, 262, 264,<br />

266, 271, 274<br />

– radicicola 116<br />

Cucumber anthracnose 36<br />

Cucumber beetles 35<br />

Cucumber mosaic virus 36<br />

Cucumis sativus 302<br />

Cucurbit 210, 217<br />

Culmorin 146<br />

Cultivation-independent methods 311<br />

Culture extracts 266<br />

Culture morphology 120<br />

Curtobacterium 91, 155<br />

– albidum 56<br />

– flaccumfaciens 23, 56<br />

– luteum 56<br />

Curvularia 164, 165, 271, 274<br />

– cymbopogonis 165<br />

– pallescens 166


358 Subject Index<br />

Cyanobacteria 155<br />

Cyclic β-glucans 73<br />

Cylindrocarpon 181<br />

– aquaticum 184<br />

– destructans 159, 181<br />

– didymum 116, 123<br />

Cypripedium calceolus 167<br />

Cystopage 193<br />

Cytogloeum sp. 166<br />

Cytokinin 269<br />

Dactylaria sp. 164<br />

Dactylorhiza majalis 167<br />

Dark septate endophytes 1, 108, 109, 114,<br />

117–122, 125, 126, 207, 209, 254, 262, 265,<br />

266, 268, 270, 272, 274, 290<br />

Dark septate fungi 290<br />

Decision-making 108<br />

Defence metabolites 145, 271<br />

Defence responses 8<br />

Definition of fungal endophytes 134<br />

Degenerate primers 341<br />

Deleterious 346<br />

Denaturing gradient gel<br />

electrophoresis 341<br />

Dendrobium moschatum 155<br />

Deoxynivalenol 144<br />

Detection 5<br />

Detrimental 340<br />

Deuteromycetes 192<br />

Diacetoxyscripenol 146<br />

Diaporthales 210<br />

Diaporthe 210, 218<br />

Differential interference<br />

microscopy 324<br />

Differential labelling 329<br />

Dikaryomycota 180<br />

Diplazium<br />

– esculentum 182, 185, 186<br />

Discosoma 326<br />

Disease 179<br />

Disease incidence 61–63<br />

Disease severity 61–63<br />

Distinguish endophytic from mycorrhizal<br />

associations 285<br />

Disturbances 108, 117<br />

– anthropogenic 108, 117, 125<br />

– natural 108, 117, 121<br />

Diversity 5, 337<br />

– fungal 180<br />

Diversity indices 21<br />

DNA 344<br />

DON 144<br />

Dothideomycetidae 210<br />

Douglas-fir (Pseudotsuga menziesii) 98<br />

Drechmeria coniospora 193, 194<br />

Drechslera australensis 163<br />

– ellisii 163<br />

Dryas octopetala 109, 110<br />

DSE 1, 108, 109, 114, 117–122, 125, 126,<br />

207, 209, 254, 262, 265, 266, 268, 270, 272,<br />

274, 290<br />

DSM 207<br />

Dual ECM/VAM associations 286<br />

Ear rots 133<br />

ECM 109<br />

Ecosystem 108, 109<br />

– forest 109, 112, 113, 117, 121, 123, 124<br />

– pristine 108<br />

Ectendomycorrhiza 4, 214, 215, 263, 286<br />

Ectomycorrhiza 209, 215–217, 263, 266<br />

– defining 286<br />

– intermediate associations 286<br />

Ectorhizosphere 195<br />

Egg- and female parasitic fungi 192<br />

Eggplant 112<br />

Elm (Ulmus) 91<br />

Endoparasitic fungi 192<br />

Endophyte 1<br />

– dark septate 158<br />

– roles 282<br />

Endophytic<br />

– definition 281<br />

Endophytic phase 283<br />

Enniatins 146<br />

Enterobacter 91<br />

– amnigenus 27<br />

– asburiae 24, 40<br />

– cloacae 27<br />

Environment 109, 339<br />

– alpine 109, 119–121<br />

– arctic 109<br />

– temperate 109, 118, 120, 121<br />

– tropical 109<br />

Environmental factors 107


Subject Index 359<br />

Epacrid mycorrhiza 212<br />

Epacridaceae 247<br />

Epacris pulchella 250<br />

Epichloë 125, 134<br />

Epicoccum 169<br />

– andropogonis 162–165<br />

– nigrum 163, 165<br />

Epiphytes 154<br />

Epulorhiza, Moniliopsis 166<br />

Equisetin 146<br />

Ergot alkaloids 134<br />

Erica arborea 209, 254<br />

– carnea 109, 110<br />

Ericaceae 212, 227<br />

Ericales 234<br />

Ericoid mycorrhiza 4, 209, 212,<br />

262, 290<br />

Erwinia carotovora 24<br />

– herbicola 330<br />

– persicinus 56<br />

– tracheiphila 34<br />

Erythromyces crocicreas 161<br />

Ethanol 301<br />

European beech 116<br />

Exoenzymes 7<br />

Exploitation 6<br />

Exploitive 1<br />

Extinctions 124<br />

Extracellular enzymes 255<br />

Extracellular polysaccharides 73<br />

Extraction 341<br />

Fagus sylvatica 116, 304<br />

Favolaschia thwaitesii 159<br />

Fen 229<br />

Ferns 183<br />

Fertilisation 25<br />

Fertilisers 108, 117, 118<br />

Filosporella 182, 184<br />

– fistucella 184<br />

– versimorpha 184<br />

Fingerprinting 337<br />

Fire 211<br />

Flagella assembly 77<br />

Flagellospora<br />

– curvula 184<br />

– fusarioides 184<br />

– penicillioides 184<br />

Flavobacterium 155<br />

Flavonoids 72<br />

Flemingia congesta 81<br />

Flexibacter 100<br />

Flooding 115<br />

Fluorescence microscopy 325<br />

Fluorescent in situ hybridisation 340<br />

Fomes 168<br />

Fomitopsis pinicola 231<br />

Fontanospora fusiramosa 182, 184<br />

Food safety 27<br />

Forest-management 117<br />

Foundations 124<br />

Founder genome 125<br />

Fumonisins 144<br />

Fungal plant pathogen 58<br />

Fungi<br />

– ancestral 180, 187<br />

– Ascomycetes 169<br />

– Basidiomycetes 169<br />

– biomass 182<br />

– clavicipitaceous 179<br />

– Coelomycetes 169<br />

– Deuteromycetes 169<br />

– foliicolous 187<br />

– Hyphomycetes 169<br />

– lignicolous 187<br />

– mycorrhizal 153<br />

– phyllosphere 179<br />

Fungi in roots<br />

– identification 291<br />

Fungicide 118, 172<br />

Fusamarin 146<br />

Fusaric acid 145<br />

Fusarium 116, 133, 163, 167, 181,<br />

262–264, 270, 271, 287<br />

– culmorum 140<br />

– graminearum 133, 139, 140, 144,<br />

145, 147<br />

– moniliforme 4, 133, 268, 274, 276<br />

– nivale 140<br />

– oxysporum 140, 160, 162, 163, 166<br />

– oxysporum f. sp. melonis 140<br />

– oxysporum f. sp. pisi 41<br />

– oxysporum f. sp. radicis lycopersici<br />

41, 324<br />

– oxysporum f. sp. vasinfectum 40<br />

– sambucinum 164


360 Subject Index<br />

– verticillioides 4, 7, 133, 136, 139, 140,<br />

144, 145, 147<br />

Fusarium wilt 40<br />

Fusarochromanone 146<br />

Fusiform rust 38<br />

Fusoproliferatum 146<br />

Gaeumannomyces graminis 116, 120, 198<br />

Ganoderma 168<br />

– australe 161<br />

Gap formation 118<br />

Gaultheria shallon 212, 237, 249, 250, 260<br />

GC clamp 342<br />

Gelatinosporium spp. 166<br />

Genetic diversity 121, 122, 125<br />

Geniculospora 184<br />

Genotype 339<br />

Geographical isolation 124<br />

Geography 22, 108<br />

Geotrichopsis sp. 164<br />

Germination<br />

– seed 156<br />

Gibberella fujikuroi 134<br />

Glaciations 124<br />

Gliocladium roseum 166<br />

Globodera pallida 194<br />

Glomerella cingulata 162, 163, 165, 166<br />

Glomeromycota 180<br />

– endophytic phases 283<br />

Gluconacetobacter diazotrophicus 338<br />

Glume blotch 116<br />

Gossypium hirsutum 302<br />

Grass 120, 125<br />

Green fluorescent protein 326<br />

Green kuang futsoi 38<br />

Gremmeniella abietina 121<br />

Group-specific PCR 346<br />

Growth enhancement 266, 274<br />

Growth stage 339<br />

Guignardia 162, 163, 170<br />

Gymnascella dankalienses 236<br />

Gyoerffyella 188<br />

Hadrotrichum 162–165, 169<br />

Hair roots 247<br />

Halep pine 209<br />

Hartig net 119<br />

Heavy metals 113, 118<br />

Helicobacter pylori 172<br />

Heliscus lugdunensis 184, 187<br />

Helotiales 212, 253<br />

Hemibiotrophic 138, 147<br />

Herbaspirillum 321<br />

– seropedicae 338<br />

Herbicide 180<br />

Herbivores 107, 179<br />

Heterobasidion annosum 230, 267<br />

Heteroconium chaetospira 4, 13, 262,<br />

270, 274<br />

Hevea<br />

– brasiliensis 182, 185<br />

Hexalectris spicata 167<br />

Hirsutella rhossiliensis 198–200<br />

Histology 262<br />

Hohenbuehelia 192, 198<br />

Holm oak 209<br />

Hordeum vulgare 303<br />

Horizontal transmission 5<br />

Host adaptation 268<br />

Host defence 116, 270<br />

Host preference 5, 116, 117<br />

Host specificity 116<br />

HrpF 81<br />

Human pathogens 26, 346<br />

Humicola 163, 164<br />

– fuscoatra 159<br />

Hyaline hyphae 263<br />

Hydrogen peroxide 301<br />

Hymenochaete crocicreas 161<br />

Hymenoscyphus ericae 212, 216, 249,<br />

253, 258, 290<br />

Hyphal coils 192<br />

Hyphomycetes 115<br />

– aquatic 115, 180<br />

Hypoxylon cf. unitum 163–165<br />

IAA 267<br />

Immunogold labeling 40<br />

Index of association 123<br />

Indole acetic acid 267<br />

Induced resistance 6, 25, 33, 269<br />

Infection asymptomatic 179<br />

Infection threads 71<br />

Inhibition 117<br />

Insect pests 107<br />

Interaction 24, 108, 115, 116, 126<br />

– among microorganisms 108


Subject Index 361<br />

– host-endophyte 180<br />

– hyphal 116<br />

– multitrophic 108, 115<br />

Inter-glacial periods 124<br />

ITS (Internal transcribed spacer)<br />

120–122, 208, 220<br />

ISSR (Inter-simple sequence repeat) PCR<br />

122, 123, 125, 251<br />

Interspecific hybrids 125<br />

Isolation procedures 299<br />

Isozyme 121<br />

ITS-RFLP 209<br />

ITS-RFLP analysis 251<br />

Kaskaskia sp. 166<br />

Kingella kingae 56<br />

Kitasatosporia cystargenia 56<br />

Klebsiella pneumoniae 338<br />

Kobresia myosuroides 212<br />

Lactic acid bacteria 346<br />

Larix decidua 262<br />

Lasiodiplodia 169<br />

– theobromae 162–166<br />

Lasmeniella sp. 163<br />

Late blight 37<br />

Latent pathogen 2, 218<br />

Leafage killers 117<br />

Lecanicillium lecanii 192<br />

Lentinula edodes 161<br />

Lenzites betulinus 161<br />

Leotiales 208<br />

Lepanthes 170<br />

Leptodontidium 4<br />

– orchidicola 11, 159, 160, 171, 211, 262<br />

Leptosphaerulina australis 164, 165<br />

Leucaena leucocephala 76<br />

Life history strategies 1<br />

Lignin 270<br />

Lipo-polysaccharides 73<br />

Live oak (Quercus fusiformis) 91<br />

Loam 114<br />

Loblolly pine 38<br />

Lodgepole pine (Pinus contorta Dougl. var.<br />

latifolia) 90<br />

Logging 108<br />

Lolium perenne 118, 120, 303<br />

Long cayenne pepper 38<br />

Lophiostomataceae 210<br />

Lotus japonicus 78<br />

Loweporus tephroporus 161<br />

Lucerne 211<br />

Lunulospora curvula 184<br />

Lycoperdon perlatum 161<br />

Lycopersicon esculentum 303<br />

Lyophyllum shimeji 161<br />

Lysis 53, 58, 59<br />

Maceration 306<br />

Macrothelypteris torresiana 182, 185, 186<br />

Maize 111<br />

Maize take-all fungi 120<br />

Malbranchea sp. 163<br />

Mangifera indica 182–186<br />

Mangrove 115<br />

Marasmius coniatus 161<br />

Massarina aquatica 186<br />

– walkeri 210<br />

Mechanical defence reactions 2<br />

Media 309<br />

Medicago sativa 302<br />

Mediterranean ecosystems 208<br />

Mediterranean heather 209<br />

Melanconium sp. 166<br />

Melanotus alpiniae 164<br />

Meloidogyne incognita 24, 196, 264<br />

Melon 112<br />

Mercuric chloride 301<br />

Mesorhizobium loti 77<br />

Metabolic activity 330<br />

Metabolite 109, 114, 116, 117, 121, 264<br />

– novel 187<br />

– inhibitory 116<br />

– secondary 109, 117, 121, 125, 180<br />

Metagenome 348<br />

Methodology 21<br />

Microascus cinereus 166<br />

Micrococcus varians 56<br />

Microcyclus sp. 166<br />

Microdochium bolleyi 114, 116, 120<br />

Microporus affinus 161<br />

Microsclerotia 118, 119, 125, 266<br />

Microspermy 241<br />

Minerals 114<br />

Mitochondria 345<br />

MLH 123–125<br />

Molecular techniques 337


362 Subject Index<br />

Monacrosporium ellipsosporum 196<br />

– haptotylum 193<br />

Moniliopsis 166<br />

Monosporascus cannonballus 216<br />

Monotropes 238<br />

Morchella 286<br />

Morphospecies 157<br />

Mortierella elongata 270, 274<br />

MRA 119<br />

Mucor 164, 165<br />

Multilocus haplotypes 123, 124<br />

Musa acuminata 302<br />

Mutation 125<br />

– rate 109, 116, 125<br />

– somatic 125<br />

Mutualism 6, 155<br />

Mutualistic root symbioses 261<br />

Mutualistic symbionts 338<br />

Mycelium radicis atrovirens 119, 166, 290<br />

Mycena orchidicola 159, 171<br />

– osmundicola 161<br />

– thuja 160<br />

Mycobacterium 155<br />

Mycobiont 235<br />

Mycocentrospora 185<br />

– acerina 181<br />

– clavata 185<br />

– iqbalii 185<br />

Mycoheterotroph 4, 238<br />

Mycoparasitism 195<br />

Mycorrhiza 116, 117, 153<br />

– arbuscular 109, 119<br />

– definition 281<br />

– ecto- 109<br />

– ericoid 229<br />

– evolution 292<br />

– orchid 238<br />

Mycorrhizal fungi 1, 25<br />

– designation 292<br />

– endophytic phase 283<br />

– inoculum 283<br />

– roles 282<br />

– structures 262<br />

Mycorrhizin 264<br />

Mycotoxin 7, 143, 144, 268, 275<br />

Myriogenospora 134<br />

Myxotrichaceae 227<br />

Myxotrichum setosum 236<br />

Nectria alata 162, 165<br />

– haematococca 162–165<br />

– ochroleuca 162, 163, 165<br />

– peziza 164<br />

– radicicola 165<br />

Nematoctonus 192, 198<br />

– pachysporus 198, 199<br />

– robustus 198–200<br />

Nematode 117<br />

– entomopathogenic 117<br />

Nematode-trapping fungi 192<br />

Nematophagous fungi 6, 191<br />

Neoplaconema napelli 165<br />

Neottia nidus-avis 167<br />

Neotyphodium 134<br />

Nicotiana tabacum 81<br />

Nigrospora sphaerica 164<br />

Nitrogen 118<br />

– fixation 90, 155<br />

– uptake 265<br />

Nitrogen-fixing bacteria 25<br />

Nivalenol 146<br />

NOx 118<br />

Nocardia 155<br />

nod-box 72<br />

NodD 72<br />

Nod-factors 72<br />

NodO 75<br />

Nodulisporium 162, 163, 165, 166<br />

– gregarium 166<br />

Non-cultivable 311, 341<br />

Non-invasive 326<br />

NopL 81<br />

NopX 80<br />

Norway spruce 114, 121–124<br />

Nostoc 155<br />

Nucleic acids 344<br />

Ochrobactrum anthropi 27<br />

Oidiodendron 249<br />

– cerealis 236<br />

– chlamydosporicum 236<br />

– citrinum 236<br />

– flavum 236<br />

– griseum 236<br />

– maius 2, 4, 227, 262, 266,<br />

274, 290<br />

– periconioides 236


Subject Index 363<br />

– rhodogenum 236<br />

– scytaloides 232, 236<br />

Oliveonia 166<br />

Optical trapping microscopy 331<br />

Optimum carrying capacity 16<br />

Orbilia 193<br />

Orchid 4, 153, 216<br />

Orchid mycorrhizas 287<br />

– designation 287<br />

Orchidaceae 153<br />

– evolution 293<br />

Ordination 179<br />

Organic amendment 26<br />

Organicization 231<br />

Oryza sativa 304<br />

Oscillatoria 155<br />

Oxytropis campestris 303<br />

Pachyrhizus tuberosus 78<br />

Paecilomyces 164, 165<br />

Paenibacillus 93<br />

– polymyxa 96<br />

Pantoea agglomerans 56, 349<br />

– anantis 56<br />

Papillae 197<br />

Parasitic 1<br />

Parasponia andersonii 71<br />

Pathogen 53–55, 57–60, 67, 107, 116,<br />

121, 179<br />

PCR 121–123, 157<br />

PCR-DGGE 341<br />

PCR-single strand conformational<br />

polymorphism (SSCP) 341<br />

PCR-TGGE 341<br />

PCR-T-RFLP 341<br />

Pea 41<br />

Peat 227<br />

Peatlands 227<br />

Pelotons 154<br />

Penicillium 163, 170<br />

– thomii 232<br />

Periconia macrospinosa 114<br />

Periconiella sp. 162, 165<br />

Peronospora parasitica 43<br />

– tabacina 37<br />

Pestalotia 163, 164, 169<br />

– adusta 166<br />

– cf. heterocornis 163, 164<br />

– poppola 163<br />

Pestalotiopsis 171<br />

– aquatica 163, 165<br />

– gracilis 163<br />

– papposa 162–164<br />

Pesticides 108<br />

Pezicula 8, 11, 13<br />

Pezizales 208<br />

Pezizella ericae 236<br />

pH 110, 114<br />

Phaeocryptopus gaeumannii 313<br />

Phaeoseptoria cf. vermiformis 163<br />

Phalangispora constricta 185<br />

Phase-contrast microscopy 324<br />

Phaseolus vulgaris 303<br />

Phellinus 168<br />

Phenolic metabolites 271<br />

Phenotypic plasticity 263<br />

Phenylpropanoids 271<br />

Pheromones 117<br />

Phialospora sp. 166<br />

Phialocephala 158, 211, 262<br />

– compacta 121<br />

– dimorphospora 121<br />

– fortinii 2, 11, 108, 109, 114–126, 159,<br />

211, 214, 217, 240, 250, 252, 254, 262, 266,<br />

270, 274, 290<br />

– scopiformis 121<br />

Phialophora 120, 211, 262<br />

– finlandica 120, 249<br />

– graminicola 120<br />

– radicicola 120<br />

– zeicola 120<br />

Phloem 338<br />

Phoma 163, 164, 166, 265<br />

Phomatospora berkeleyi 166<br />

Phomopsis 163, 169, 210, 218<br />

– cf. orchidophila 162–166<br />

Phosphorus 114, 265, 266, 269<br />

Photosynthates 109<br />

Phragmites australis 111, 115<br />

Phyllobacterium 93<br />

– rubiacearum 61<br />

Phyllosticta capitalensis 166<br />

– colocasiicola 166<br />

Physiological relationships 143<br />

Phytoalexin 171


364 Subject Index<br />

Phytohormones 7, 267, 268, 273<br />

Phytophthora cactorum 55<br />

– cinnamomi 230<br />

– infestans 37<br />

Phytoplasms 338<br />

Phytosphere 348<br />

Phytosterol 270<br />

Picea abies 110, 114, 119, 121–124,<br />

303, 304<br />

– glauca 118, 182–186<br />

Piceirhiza bicolorata 212<br />

Pili 77<br />

Pinus contorta 302<br />

– halepensis 209<br />

– muricata 118<br />

– sylvestris 112, 116, 119, 212, 216, 304<br />

– taeda 38<br />

Piriformospora indica 4, 13, 201, 262,<br />

267, 272<br />

Pisum sativum 75<br />

Pithomyces maydicus 162–165<br />

Plant and Fungus Interactions 134<br />

Plant defence mechanisms 25<br />

Plant disease 65, 67<br />

Plant genotype 23<br />

Plant growth 53, 58, 60, 62, 66, 68<br />

Plant health 53, 345<br />

Plant pathogens 24<br />

Plant species 22<br />

Plant symbionts 25<br />

Plant-associated 338<br />

Plant-fungus coevolution 282<br />

Plant-metagenomic approach 348<br />

Plants<br />

– myco-heterotrophic 154<br />

Plasmodiophora brassicae 270<br />

Plasticity 9, 269, 273<br />

Plectosporium tabacinum 59, 61<br />

Pleosporales 208, 210<br />

Pleurotus djamor 198, 199<br />

– ostreatus 192<br />

Pochonia 193<br />

– bulbillosa 232<br />

– chlamydosporia 193, 197, 198<br />

– rubescens 193<br />

Pollution 108, 113, 118<br />

Polymerase chain reaction 121, 123, 340<br />

Polyphasic 348<br />

Population densities 15<br />

Population genetics 119, 122<br />

Populations 108, 121, 123, 338<br />

Populus hybrida 182–186<br />

Potato (Solanum tuberosum) 90, 111, 117<br />

Pratylenchus 201<br />

Preceding crop 117<br />

Pressure bomb method 21, 309<br />

Pressure extraction 308<br />

Primers 341<br />

Proteolytic enzymes 194<br />

PR-proteins 81<br />

Pseudallescheria boydii 165<br />

Pseudoanguillospora 185<br />

Pseudogymnoascus roseus 236<br />

Pseudomonas 54, 56, 60, 66–69, 91,<br />

155, 321<br />

– chlororaphis 61, 94<br />

– chlororaphis PCL1391 327<br />

– cichorii 23, 56<br />

– corrugata 56<br />

– denitrificans 91<br />

– fluorescens 22, 34, 56, 61, 65, 66,<br />

68, 146<br />

– fluorescens CHA0 329<br />

– fluorescens SBW25 330<br />

– fluorescens WCS365 322<br />

– graminis 56, 61<br />

– migulae 56<br />

– orientalis 56<br />

– putida 22, 56, 61, 91, 330<br />

– reactans 56<br />

– rhodesiae 56<br />

– straminea 56<br />

– synthaxa 56<br />

– syringae 56, 90<br />

– syringae pv. lachrymans 34<br />

– tolaasii 56, 61<br />

– trivialis 61<br />

– veronii 56, 61<br />

Pseudomycorrhiza 119<br />

Pseudotsuga menziesii 117<br />

Psychilis kraenzlinii 170<br />

Psychrobacter immobilis 56<br />

Pteridium aquilinum 115, 303<br />

Pterostylis 155<br />

Purification 341<br />

Pyrenochaeta cf. rubi-idaei 164


Subject Index 365<br />

Pythium 268<br />

– oligandrum 271, 274<br />

– spinosum 55<br />

Quantification 5, 313<br />

Quercus ilex 209, 255<br />

Quorum sensing 324<br />

Ralstonia paucula 56<br />

– picketti 346<br />

– solanacearum 38, 340<br />

Ramichloridium apiculatum 166<br />

– cf. subulatum 164<br />

RAPD (Random amplified polymorphic<br />

DNA) 122, 214<br />

rDNA 208, 219<br />

Real-time PCR 340<br />

Recombination 122, 125<br />

Red fluorescent protein 326<br />

Reporter genes 322<br />

Resistance 116<br />

Restriction fragment length<br />

polymorphism 121<br />

Restriction patterns 120<br />

RFLP 121, 122, 125<br />

Rhizobacteria 34, 201<br />

Rhizobiaceae 321<br />

Rhizobium etli 25<br />

– leguminosarum 25, 75<br />

– meliloti 56, 71, 76<br />

– sp. NGR234 71<br />

Rhizoctonia 4, 156, 166, 287<br />

– solani 24, 38, 55, 59, 61, 195<br />

Rhizomonas suberifaciens 56<br />

Rhizophora mucronata 110, 115, 183, 186<br />

Rhizoplane 107<br />

Rhizopycnis vagum 210<br />

Rhizoscyphus ericae 231<br />

Rhizosphere 107, 114, 117, 125, 322<br />

Rhodococcus 155<br />

Ribosomal 340<br />

Ribosomal RNA 120<br />

Rice (Oryza sativa) 90, 120<br />

– crown sheath rot 120<br />

RNA 343<br />

rRNA 120, 121<br />

Root endophytes, multiple roles 290<br />

Root nodules 71<br />

Root turnover 109<br />

Root-colonising 321<br />

Roots, adventitious 155<br />

Rosemary 209<br />

Rosmarinus officinalis 209<br />

Rot 133<br />

Rough lemon (Citrus jambhiri) 90<br />

16S rDNA 341<br />

Salicylic acid 34<br />

Salix babylonica 182–186<br />

Salmonella enterica 27<br />

Saprobe 2, 228<br />

Saprophytic 266<br />

Saprotrophic symbiont 219<br />

Scanning electron microscopy 324<br />

Sclerroderis canker 121<br />

Scots pine 116<br />

Scytalidium vaccinii 236<br />

Seawater 115<br />

Sebacina 156, 166, 238, 250<br />

sec pathway 78<br />

Secondary metabolites 264, 273<br />

Sedges 120<br />

Seed borne 137<br />

Seed treatment 118<br />

Selective media 311<br />

Septoria nodorum 116, 118<br />

Sequences, ribosomal 180<br />

Serendipita 166<br />

Serine protease 194<br />

Serratia 54, 56<br />

– marcescens 26, 35, 65, 66, 338, 346<br />

– plymuthica 58, 61, 66, 67<br />

Sink 269, 270, 272<br />

Sinorhizobium meliloti 329<br />

Ski slopes 118<br />

SO2 118<br />

Sodium hypochlorite 301<br />

Soil suppressiveness 202<br />

Soil texture 114<br />

Soil-rhizosphere-root continuum 107<br />

Solanum tuberosum 90, 111, 117, 302<br />

Sonification 300<br />

Sonneratia caseolaris 183, 186<br />

Sordaria fimicola 159<br />

Sordariomycetidae 210<br />

Spanish oak (Quercus texana) 91<br />

Spatiotemporal patterns 108


366 Subject Index<br />

Speciation 123, 125, 126<br />

– allopatric 123, 125<br />

– sympatric 123<br />

Species 110–113, 345<br />

– abundance 108, 115, 121<br />

– competitive 107, 116, 118<br />

– cryptic 123–126<br />

– diversity 108–110, 114–119, 121–126<br />

– evenness 108<br />

– identification 120<br />

– richness 108, 109, 114, 115, 117, 118<br />

– spectrum 108, 109, 115–117<br />

– typing 120<br />

Specificity 5, 158<br />

Sphagnum 229<br />

Sphigobacterium thalophilum 56<br />

Sphingomonas 91<br />

– adhaesiva 56<br />

– trueperi 61<br />

Spore production 182<br />

Sporormia minima 160<br />

Spruce 229<br />

Stagonospora 263, 265, 274<br />

Staphylococcus xylosus 27<br />

Stem rot 133<br />

Stenotrophomonas 91<br />

– maltophilia 26, 56<br />

Sterilising agents 301<br />

Sterility check 300, 305<br />

Stimulation 117<br />

Strain mixtures 38<br />

Strawberry 112<br />

Streptomyces 93, 321<br />

– bottropensis 56<br />

– diastatochromogenes 56<br />

– galilaeus 56<br />

– griseus 56<br />

– lavendulae 56<br />

– scabies 55<br />

– setonii 56<br />

– turgidiscabies 56<br />

Stress tolerance 274<br />

Stylopage 193<br />

Succession 122, 124<br />

Sugar beet 111<br />

Sugarcane Saccharum officinarum 90<br />

Suppression 346<br />

Surface sterilisation 5, 300<br />

Surfactants 305<br />

Symbiont 235<br />

Symbiosis 1<br />

Sympatric occurrence 123<br />

Symptomless root infections 139<br />

Systemic acquired resistance 33<br />

Systemic fungal root infections 269<br />

Systemic resistance 271<br />

T-2 toxins 146<br />

Take-all fungus 198<br />

Taxon 344<br />

Tecomella undulata (Bignoniaceae) 90<br />

Teleomorph 181, 186, 187<br />

Temperature gradient gel<br />

electrophoresis 341<br />

Tephrosia vogelii 78<br />

Terminal restriction fragment length<br />

polymorphism 341<br />

Tetrabrachium<br />

– elegans 185<br />

Tetracladium 185<br />

– furcatum 185<br />

– marchalianum 181, 185<br />

– setigerum 185<br />

Tetrazolium 348<br />

Thailand 38<br />

Thanatephorus 156, 166<br />

– gardneri 287<br />

Thermomyces verrucosus 159<br />

Thielavia 249, 253<br />

– basicola 160<br />

Thuja occidentalis 117<br />

Tidal level 115<br />

Tillandsia 155<br />

Tissue type 339<br />

Tobacco 36<br />

Tolumnia variegata 169<br />

Tomato 36, 37, 112, 210<br />

Tomato mottle virus 37<br />

Total community nucleic acids 341<br />

Total microflora 324<br />

Toxicity 179<br />

Toxin-producing fungi 192<br />

Toxins 264<br />

Trametes hirsuta 161<br />

Trapping organs 192<br />

Tricellula aquatica 185


Subject Index 367<br />

Trichocladium opacum 160<br />

Trichoderma 163, 168, 195<br />

Tricholoma 286<br />

Trichosporiella multisporum 159<br />

Tricladium chaetocladium 185<br />

– splendens 181<br />

Triscelophorus<br />

– acuminatus 185<br />

– konajensis 186<br />

– monosporus 186<br />

Triticum aestivum 111, 114, 117, 302<br />

Troposporella sp. 162–164<br />

Tsuga 212<br />

Tubercularia 164<br />

Tuberculate mycorrhizae 98<br />

Tulasnella 156, 166<br />

Tumularia aquatica 186<br />

Type 1 121<br />

Type I secretion systems 73<br />

Type II secretion systems 77<br />

Urban development 108<br />

Vaccinium corymbosum 253<br />

– macrocarpon 253<br />

– myrtillus 216<br />

Vacuum 308<br />

Valsaceae 210<br />

Vanilla 171<br />

Varicosporium 188<br />

– elodeae 186<br />

– giganteum 186<br />

Vascular cylinder 2<br />

Vascular tissue 309<br />

Velamen 154<br />

Verticillium 193, 287<br />

– chlamydosporium 264, 275<br />

– dahliae 55, 59<br />

– lecanii 166<br />

– longisporum 24, 55, 270<br />

Vicia sativa 75<br />

Vigna unguiculata 78<br />

Vine decline 210, 217<br />

Virulence factors 7, 264<br />

Virulent pathogens 2<br />

Visoltricin 146<br />

Vitis vinifera 329<br />

Volatile organic compounds 44<br />

Vomitoxin 144<br />

Water regime 114, 115<br />

Weather 109<br />

Weevil larvae 117<br />

Westerdykella 270, 274<br />

Western red cedar (Thuja plicata) 100<br />

Wetland 115<br />

Wheat (Triticum aestivum) 96, 114,<br />

116–118<br />

White x Engelmann hybrid spruce (Picea<br />

glauca x P. enelmannii 92<br />

Whitefly 37<br />

Wildfire 36, 118<br />

Windthrow 112, 117, 118<br />

Woodland 115<br />

Woollsia pungens 212, 249, 251, 253,<br />

258–260<br />

Wortmannin 146<br />

Wullschlaegelia calcarata 168<br />

Xanthomonas 155<br />

– campestris 55, 56<br />

– campestris pv. vesicatoria 81<br />

– oryzae 56<br />

Xylaria 162, 163, 164, 168<br />

– arbuscula 163<br />

– cf. cubensis 163<br />

– cf. curta 163<br />

– corniformis 163<br />

– enteroleuca 163<br />

– hypoxylon 163<br />

– mellisii 163<br />

– multiplex 163<br />

– obovata 163<br />

– polymorpha 163<br />

Xylella fastidiosa 313, 329<br />

Xylem 338<br />

Yield shield 44<br />

Ypsilonidium 166<br />

Zea mays 302<br />

Zearalenone 144<br />

Zygomycetes 193

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!