Academia.eduAcademia.edu
Molecular Identification of Fungi Youssuf Gherbawy Kerstin Voigt Editors 123 Molecular Identification of Fungi . Youssuf Gherbawy l Kerstin Voigt Editors Molecular Identification of Fungi Editors Prof. Dr. Youssuf Gherbawy South Valley University Faculty of Science Department of Botany 83523 Qena, Egypt youssuf_gherbawy@hotmail.com Dr. Kerstin Voigt University of Jena School of Biology and Pharmacy Institute of Microbiology Neugasse 25 07743 Jena, Germany kerstin.voigt@uni-jena.de ISBN 978-3-642-05041-1 e-ISBN 978-3-642-05042-8 DOI 10.1007/978-3-642-05042-8 Springer Heidelberg Dordrecht London New York Library of Congress Control Number: 2009938949 # Springer-Verlag Berlin Heidelberg 2010 This work is subject to copyright. All rights are reserved, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965, in its current version, and permission for use must always be obtained from Springer. Violations are liable to prosecution under the German Copyright Law. The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. Cover design: WMXDesign GmbH, Heidelberg, Germany, kindly supported by ‘leography.com’ Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com) Dedicated to Prof. Lajos Ferenczy (1930–2004) microbiologist, mycologist and member of the Hungarian Academy of Sciences, one of the most outstanding Hungarian biologists of the twentieth century . Preface Fungi comprise a vast variety of microorganisms and are numerically among the most abundant eukaryotes on Earth’s biosphere. They enjoy great popularity in pharmaceutical, agricultural, and biotechnological applications. Recent advances in the decipherment of whole fungal genomes promise a continuation and acceleration of these trends. New techniques become available to facilitate the genetic manipulation of an increasing number of fungal organisms to satisfy the demand of industrial purposes. The increasing importance-driven search of novel detection techniques and new fungal species initiated the idea for a book about the molecular identification of fungi. The kingdom of the fungi (Mycota) appears as the sister group of the multicellular animals (Metazoa) as an independent, apparently monophyletic group within the domain Eukarya, equal in rank to green plants (Viridiplantae) and animals (Metazoa). Fungi are originally heterotrophic eukaryotic microorganisms harboring chitin in their cell walls and lacking plastids in their cytoplasm. Formerly, the oomycetes, slime moulds and plasmodiophorids were considered as fungi based on their ability to produce fungus-like hyphae or resting spores. Whereas the Oomycota are classified to the stramenopile algae (Chromista or Heterokonta), and the plasmodial and cellular slime moulds (Mycetozoa) belong to the Amoebozoa. The Plasmodiophoromycota are among the cercozoan Rhizaria closely related to the foraminifers. A three-protein phylogeny of the fungi and their allies confirms that the nucleariids, phagotrophic amoebae with filose pseudopods in soil and freshwater, may represent descendants of a common ancestor at the animal–fungal boundary (Fig. 1). The fungal kingdom encompasses the Asco-, Basidio-, Glomero-, Zygo- and Chytridiomycota. The former four phyla are terrestrial fungi developing nonflagellated spores (aplanosporic), whereas the Chytridiomycota represent aquatic and zoosporic (planosporic) fungi, which split into three individual taxon groups, the aerobic Blastocladio- and Chytridiomycota sensu stricto and the anaerobic Neocallimastigomycota. The Zygomycota are among the most basal terrestrial fungi, which evolved in a paraphyletic manner. Hence, the phylum was divided into different subphyla, vii viii Preface Fig. 1 The evolution of the fungi and allied fungi-like microorganisms based on a concatenated neighbor-joining analysis using mean character differences as distance measure on 1,262 aligned amino acid characters comprising translation elongation factor 1 alpha, actin, and beta-tubulin (500, 323 and 439 characters, respectively) from 80 taxa. The prokaryotic elongation factor Tu, MreB (TM1544), and FtsZ (both homologous to actin and tubulin, respectively) from Thermotoga maritima were used as out group taxon representing the bacterial domain the Mucoro-, Kickxello-, Zoopago- and Entomophthoromycotina, whose phylogenetic relationships are not fully understood yet. In the phylogenetic tree shown in Fig. 1, the Entomophthoromycotina group together with the Ichthyosporea, a relationship, is not well supported by clade stability proportions. Fungi develop a wide diversity of morphological features, which are shared with many fungi-like microorganisms (Fig. 2), among those the white rust and downy mildew “fungi” (Fig. 2g) are obligate parasites of plants and develop fungus-like hyphae with haustoria (ht) in asexual and thick-walled, ornamented oospores (os) from fertilized oospheres after fusion of an oogonium (og) with an antheridium (at) during sexual reproduction (Fig. 3). The distribution of fungi among the various ecological niches of the biosphere seems to be infinite. Estimates suggest a total of 1.5 million fungal species, only less than a half has been merely described yet. This implies a backlog demand, which comes along with a rising importance of novel techniques for a rapid and Preface ix Fig. 2 The morphological diversity of fungi and fungi-like microorganisms. (a–f ): basidiomycetes (Agaricomycotina; Photos: M. Kirchmair); (g) oomycetes (Peronosporales; Photo: O. Spring); (h–j): multicellular conidia from imperfect stages of ascomycetes (Pezizomycotina); (k–s): zygomycetes (Mucoromycotina; Photos: K. Hoffmann, scanning electron microphotographs o & q: M. Eckart & K. Hoffmann): (k, l, p, r, s) – different types of multispored sporangia, (m, n, o): different types of uni‐fewspored sporangiola; (t–x): reproductive structures (zoosporangia) from anaerobic chytridiomycetes (Neocallimastigomycota; Photos: K. Fliegerova); (y, z): plasmodiophorids (Plasmodiophoromycota; Photos: S. Neuhauser & M. Kirchmair). x Preface Fig. 3 Cross-section of a leaf infected with Pustula tragopogonis (Peronosporales, Oomycota) causing white rust on sunflower. The microphotograph shows structures, which are typical for the sexual reproduction of oomycetes: ht – haustorium, ld – lipid droplet inside an oospore, os – oospore, og – oogonium, at – antheridium fused to an oogonium (Photo: A. Heller) unambiguous detection and identification of fungi to explore the fungal diversity as a coherent whole. Molecular techniques, particularly the technology of the polymerase chain reaction, have revolutionized the molecular biology and the molecular diagnosis of fungi. The incorporation of molecular techniques into what has been traditionally considered as morphology-based taxonomy of fungi helps us in the differentiation of fungal species and varieties. Databases of genomes and genetic markers used as sources for molecular barcodes are being created and the fungal world is in progress to be unveiled with the help of bioinformatics tools. Genome projects provide evidence for ancient insertion elements, proviral or prophage remnants, and many other patches of unusual composition. Consequently, it becomes increasingly important to pinpoint genes, which characterize fungal organisms at different taxonomic levels without the necessity of previous cultivation. Unfortunately, the initiative of an excessive use of molecular barcoding has been hampered by a lack of sufficient and novel synapomorphic nucleotide < Fig. 2 (continued) (a) – basidiocarp of Schizophyllum commune, (b) – basidiocarp of Daedalea quercina, (c) – hymenophor from basidiocarp of Daedalea quercina, (d) – basidiocarp of Trametes sp., (e) – mycelium of Antrodia sp spreading over a trunk of a tree, (f ) – dry rot caused by Serpula lacrymans on timber, (g) –symptomatology from Plasmopara viticola, the causal agent of grapevine downy mildew, (h) – Pestalotiopsis clavispora (Photo: C. Kesselboth), (i) – Bipolaris cf. sorokiniana (Photo: G. Newcombe), ( j) – Fusarium sp. (Photo: C. Kesselboth), (k) – Mucor indicus, (l) – Helicostylum elegans, (m) – Thamnidium elegans, (n) – Dichotomocladium sp., (o) – Dichotomocladium robustum, (p) – Absidia psychrophilia, (q) – zygospores from Lentamyces parricida, (r) – Mucor rouxii, (s) – Absidia cylindrospora, (t) – Caecomyces sp. isolated from sheep (lugol staining), (u) – Caecomyces sp. isolated from sheep, (v) – Neocallimastix frontalis (bisbenzimide staining of nuclei), (w) – Anaeromyces mucronatus isolated from cow (bisbenzimide staining of nuclei), (x) – Neocallimastix frontalis isolated from cow (lugol staining); (y) – thick walled resting spores from Sorosphaera veronicae, (z) – sporosori from Sorosphaera veronicae Preface xi characters and signature sequences. Moreover, high intraspecific variability of conventional molecular characters makes it difficult to identify species borders. However, DNA sequences and other genetic markers provide large amounts of data which are cultivation-independent and do not depend on physiological inconsistencies. Genetic markers constantly reflect the identification treasure hidden in the genetic information and allow to control the degree of resolution by choosing the appropriate genes. In this book, we highlight the advances of the past decade, both in methodology and in the understanding of genomic organization and approach problems of the identification and differentiation of fungi using molecular markers and compare those with classical procedures traditionally used for species designation. The limitations in the availability of type material, reference strains, and reference nucleotide sequences set boundaries in the molecular identification. For example, the image displaying multicellular, melanin-pigmented conidia (size: 90 mm) from strain CID1670 (Fig. 2i), which was kindly provided by George Newcombe (University of Idaho, Center for Research on Invasive Species and Small Populations, Moscow, ID, USA), may serve as an appropriate cautionary note for readers of this book. The strain was recovered as an endophytic ascomycete from the asterid perennial herb Centaurea stoebe (spotted knapweed). The fungus could be attributed by conventional ITS barcoding to the pleosporalean genus Drechslera and in a narrower sense to Bipolaris sorokiniana. Since species of Bipolaris had never been reported from any species of Centaurea in earlier reports, neither its effects on its host nor the final taxonomic delimitation are known. Nucleotide sequences of additional genes and a more in-depth phylogenetic study may even suggest that this strain was a new species. Therefore, it would make sense to distinguish between refined identification of fungi uncommonly found in exceptional biotopes in order to explore new species, e.g., as endophytes, and high-throughput molecular identification of well-studied fungi in order to serve the needs of industrial application. The role of fungi as pathogens of evolutionarily naive plants including a hypothesis about the plant invasion-mediated progression of novel phytopathogens will be discussed in the first chapter. The second and third chapter concerns with the diagnostics and the challenge to identify “fungus-like” plant pathogens from the oomycetes and the plasmodiophorids, respectively. The fourth chapter leads over the applications of molecular markers and DNA sequences in the identification of fungal pathogens in grain legumes and cereals followed by various aspects of qualitative and quantitative detection of Fusarium spp. and Macrophomina phaseolina, pathogenic on maize and other corn crops or economic plants. During the course of the book, the detection of ochratoxigenic fungi, mainly aspergilli and penicilli, and other postharvest pathogens like Mucor and Rhizopus is elucidated. The molecular identification of wood rotting and endophytic fungi as well as anaerobic rumen fungi finish the first part on plant pathological and environmental biological aspects. The second part deals with human pathological and clinical aspects. The introduction gives a contribution about new approaches in fungal DNA preparation from whole blood following multiplex PCR detection. Novel techniques in the depletion of the background host DNA in favour of enrichment of the xii Preface fungal contaminant DNA following different modifications of PCR approaches represent powerful tools in the detection of a wide variety of human pathogenic fungi causing sepsis and other life-threatening diseases that result from excessive host responses to fungal infections. The survey continues with conventional strategies for the molecular detection of Malassezia, dermatophytes, opportunistic fungi, and causative agents of deep mycoses as well as paracoccidioidomycosis and Ochroconis gallopava infection via a novel tool, the loop-mediated isothermal amplification method (LAMP). The book closes with reviews about prospects and perspectives of molecular markers for the identification of Absidia-like fungi and other zygomycetes. The editors thank all contributors for their valuable reviews and comments, which were crucial for the accomplishment of this book. Furthermore, we express our gratitude to all authors who contributed figures and images for the cover and miscellaneous parts adding a great deal to the illustration of this book. The cover of the book was kindly supported by “leography.com.” January 2010 Youssuf Gherbawy Kerstin Voigt Contents Part I Plant Pathological and Environmental Biological Aspects 1 Fungal Pathogens of Plants in the Homogocene . . . . . . . . . . . . . . . . . . . . . . . . 3 George Newcombe and Frank M. Dugan 2 Molecular Techniques for Classification and Diagnosis of Plant Pathogenic Oomycota . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 Otmar Spring and Marco Thines 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs with a Multiphasic Life Cycle . . . . . . . . . . . . . . . . . . . . 51 Sigrid Neuhauser, Simon Bulman, and Martin Kirchmair 4 Applications of Molecular Markers and DNA Sequences in Identifying Fungal Pathogens of Cool Season Grain Legumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 Evans N. Njambere, Renuka N. Attanayake, and Weidong Chen 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 Kurt Brunner and Robert L. Mach 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 Ivan Visentin, Danila Valentino, Francesca Cardinale, and Giacomo Tamietti 7 Molecular Detection and Identification of Fusarium oxysporum . . . . . 131 Ratul Saikia and Narendra Kadoo xiii xiv Contents 8 Molecular Chemotyping of Fusarium graminearum, F. culmorum, and F. cerealis Isolates From Finland and Russia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159 Tapani Yli-Mattila and Tatiana Gagkaeva 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina: A Charcoal Rot Fungus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179 Bandamaravuri Kishore Babu, Ratul Saikia, and Dilip K Arora 10 Molecular Diagnosis of Ochratoxigenic Fungi . . . . . . . . . . . . . . . . . . . . . . . 195 Daniele Sartori, Marta Hiromi Taniwaki, Beatriz Iamanaka, and Maria Helena Pelegrinelli Fungaro 11 Molecular Barcoding of Microscopic Fungi with Emphasis on the Mucoralean Genera Mucor and Rhizopus . . . . . . . . . . . . . . . . . . . . . 213 Youssuf Gherbawy, Claudia Kesselboth, Hesham Elhariry, and Kerstin Hoffmann 12 Advances in Detection and Identification of Wood Rotting Fungi in Timber and Standing Trees . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251 Giovanni Nicolotti, Paolo Gonthier, and Fabio Guglielmo 13 Molecular Diversity and Identification of Endophytic Fungi . . . . . . . 277 Liang-Dong Guo 14 Molecular Identification of Anaerobic Rumen Fungi . . . . . . . . . . . . . . . . 297 Martin Eckart, Katerina Fliegerová, Kerstin Hoffmann, and Kerstin Voigt Part II Human Pathological and Clinical Aspects 15 New Approaches in Fungal DNA Preparation from Whole Blood and Subsequent Pathogen Detection Via Multiplex PCR . . . . 317 Roland P. H. Schmitz, Raimund Eck, and Marc Lehmann 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS and D1/D2 Regions of DNA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337 Lidia Pérez-Pérez, Manuel Pereiro, and Jaime Toribio 17 DNA-Based Detection of Human Pathogenic Fungi: Dermatophytes, Opportunists, and Causative Agents of Deep Mycoses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357 Lorenza Putignani, Silvia D’Arezzo, Maria Grazia Paglia, and Paolo Visca Contents xv 18 Applications of Loop-Mediated Isothermal Amplificaton Methods (LAMP) for Identification and Diagnosis of Mycotic Diseases: Paracoccidioidomycosis and Ochroconis gallopava infection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417 Ayako Sano and Eiko Nakagawa Itano 19 Identification of the Genus Absidia (Mucorales, Zygomycetes): A Comprehensive Taxonomic Revision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439 Kerstin Hoffmann 20 Molecular Characters of Zygomycetous Fungi . . . . . . . . . . . . . . . . . . . . . . . 461 Xiao-yong Liu and Kerstin Voigt Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 489 . Contributors Dilip K. Arora National Bureau of Agriculturally Important Microorganisms (ICAR), Mau, Uttar Pradesh 275101, India, aroradilip@yahoo.co.in Renuka N. Attanayake Department of Plant Pathology, Washington State University, Pullman, WA 99164, USA Kurt Brunner Institute of Chemical Engineering, Research Area Gene Technology and Applied Biochemistry, Gene Technology Group, Vienna University of Technology, Getreidemarkt 9, A-1060 Vienna Simon Bulman Plant & Food Research, Private Bag 4704, Christchurch, New Zealand; Bio-Protection Research Centre, Lincoln University, P.O. Box 84, 7647 Canterbury, New Zealand Francesca Cardinale DiVaPRA – Plant Pathology, University of Turin, I-10095 Grugliasco, Turin, Italy Weidong Chen Department of Plant Pathology, Washington State University, Pullman, WA 99164, USA; USDA ARS Grain Legume Genetics and Physiology Research Unit, Washington State University, Pullman, WA 99164, USA, w-chen@wsu.edu Silvia D’Arezzo National Institute for Infectious Diseases “Lazzaro Spallanzani” I.R.C.C.S., Via Portuense 292, 00149 Rome, Italy Frank M. Dugan USDA-ARS, Washington State University, Pullman, WA 99163-6402, USA Raimund Eck SIRS-Lab GmbH, Winzerlaer Str. 2, 07745 Jena, Germany xvii xviii Contributors Martin Eckart Institute of Microbiology, School of Biology and Pharmacy, University of Jena, Neugasse 25, 07743 Jena, Germany, martin.eckart@uni-jena.de Hesham Elhariry Biological Sciences Department, Faculty of Science, Taif University, P.O. Box 888 Taif, Kingdom of Saudi Arabia Katerina Fliegerová Department of Biological Basis of Food Quality and Safety, Institute of Animal Physiology and Genetics, Czech Academy of Sciences, v.v.i., Vı́deňská 1083, 14220 Prague 4, Czech Republic, fliegerova@iapg.cas.cz Tatiana Gagkaeva Laboratory of Mycology and Phytopathology, All-Russian Institute of Plant Protection (VIZR), 196608 St. Petersburg-Pushkin, Russia, t.gagkaeva@yahoo.com Youssuf Gherbawy Botany Department, Faculty of Science, South Valley University, 83523 Qena, Egypt Paolo Gonthier Di.Va.P.R.A., Department of Exploitation and Protection of the Agricultural and Forestry Resources, Plant Pathology, University of Torino, via L. da Vinci 44, I-10095 Grugliasco (TO), Italy Maria Grazia Paglia National Institute for Infectious Diseases “Lazzaro Spallanzani” I.R.C.C.S., Via Portuense 292, 00149 Rome, Italy Fabio Guglielmo Di.Va.P.R.A., Department of Exploitation and Protection of the Agricultural and Forestry Resources, Plant Pathology, University of Torino, via L. da Vinci 44, I-10095 Grugliasco (TO), Italy Liang-Dong Guo Systematic Mycology & Lichenology Laboratory, Institute of Microbiology, Chinese Academy of Sciences, Beijing 100101, China, guold@sun. im.ac.cn Kerstin Hoffmann Institute of Microbiology, School of Biology and Pharmacy, University of Jena, Neugasse 25, 07743 Jena, Germany, Hoffmann.Kerstin@ uni-jena.de Beatriz Iamanaka Departamento de Biologia Geral, Centro de Ciências Biológicas, Universidade Estadual de Londrina, Caixa Postal 6001, CEP 86051-970 Londrina-Paraná, Brazil Eiko Nakagawa Itano Department of Pathological Science, CCB, State University of Londrina, P.O. Box 6001, 86051-970 Londrina, Paraná, Brazil, itanoeiko@hotmail.com Contributors xix Narendra Kadoo PMB Group, Biochemical Sciences Division, National Chemical Laboratory, Pune 411008, Maharashtra, India, ny.kadoo@ncl.res.in Claudia Kesselboth Botany Department, Faculty of Science, South Valley University, 83523 Qena, Egypt Martin Kirchmair Institute of Microbiology, Leopold Franzens – University Innsbruck, Technikerstr. 25, 6020 Innsbruck, Austria, Martin.Kirchmair@uibk.ac.at Bandamaravuri Kishore Babu National Bureau of Agriculturally Important Microorganisms (ICAR), Mau, Uttar Pradesh 275101, India, aroradilip@yahoo. co.in; present address: Environmental Microbiology Lab, Department of Environmental Engineering, Chosun University, Gwang ju-501759, South Korea, kishore_bandam@yahoo.co.in Marc Lehmann SIRS-Lab GmbH, Winzerlaer Str. 2, 07745 Jena, Germany Xiao-yong Liu Key Laboratory of Systematic Mycology and Lichenology, Institute of Microbiology, Chinese Academy of Sciences, No. 1 Beichen West Road, Chaoyang District, Beijing 100101, P. R. China, liuxiaoyong@im.ac.cn Robert L. Mach Institute of Chemical Engineering, Research Area Gene Technology and Applied Biochemistry, Gene Technology Group, Vienna University of Technology, Getreidemarkt 9, A-1060 Vienna, Austria, rmach@mail.zserv.tuwien. ac.at Sigrid Neuhauser Institute of Microbiology, Leopold Franzens – University Innsbruck, Technikerstr. 25, 6020 Innsbruck, Austria George Newcombe Department of Forest Resources, and Center for Research on Invasive Species and Small Populations, University of Idaho, Moscow, ID 838441133, USA, georgen@uidaho.edu Giovanni Nicolotti Di.Va.P.R.A., Department of Exploitation and Protection of the Agricultural and Forestry Resources, Plant Pathology, University of Torino, via L. da Vinci 44, I-10095 Grugliasco (TO), Italy, giovanni.nicolotti@unito.it Evans N. Njambere Department of Plant Pathology, Washington State University, Pullman, WA 99164, USA Maria Helena Pelegrinelli Fungaro Departamento de Biologia Geral, Centro de Ciências Biológicas, Universidade Estadual de Londrina, Caixa Postal 6001, CEP 86051-970, Londrina-Paraná, Brazil, fungaro@uel.br xx Contributors Manuel Pereiro Department of Dermatology, Laboratory of Mycology, Faculty of Medicine, University Hospital Complex of Santiago de Compostela, C/San Francisco S/N, 15706 Santiago de Compostela, Spain Lidia Pérez-Pérez Department of Dermatology, University Hospital Complex of Vigo, C/Porriño 5, 36209 Vigo, Spain, lidiacomba@yahoo.es Lorenza Putignani Microbiology Unit, Children’s Hospital, Healthcare and Research Institute Bambino Gesù, Piazza Sant’Onofrio 4, 00165 Rome, Italy Ratul Saikia Biotechnology Division, North-East Institute of Science & Technology, Jorhat 785006, Assam, India, rsaikia19@yahoo.com Ayako Sano Medical Mycology Research Center, Chiba University, 1-8-1, Inohana, Chuo-ku, 260-8673 Chiba, Japan, aya1@faculty.chiba-u.jp Daniele Sartori Centro de Ciências Biológicas, Departamento de Biologia Geral, Universidade Estadual de Londrina, Caixa Postal 6001, CEP 86051-970, Londrina-Paraná, Brazil Roland P.H. Schmitz SIRS-Lab GmbH, Winzerlaer Str. 2, 07745 Jena, Germany, schmitz@sirs-lab.com Otmar Spring Institute of Botany, University of Hohenheim, 70593 Stuttgart, Germany, spring@uni-hohenheim.de Giacomo Tamietti DiVaPRA – Plant Pathology, University of Turin, I-10095 Grugliasco, Turin, Italy, giacomo.tamietti@unito.it Marta Hiromi Taniwaki Departamento de Biologia Geral, Centro de Ciências Biológicas, Universidade Estadual de Londrina, Caixa Postal 6001, CEP 86051970, Londrina-Paraná, Brazil Marco Thines Institute of Botany, University of Hohenheim, 70593 Stuttgart, Germany Jaime Toribio Department of Dermatology, Laboratory of Mycology, Faculty of Medicine, University Hospital Complex of Santiago de Compostela, C/San Francisco S/N, 15706 Santiago de Compostela, Spain Danila Valentino DiVaPRA – Plant Pathology, University of Turin, I-10095 Grugliasco, Turin, Italy Contributors xxi Paolo Visca National Institute for Infectious Diseases “Lazzaro Spallanzani” I.R.C.C.S., Via Portuense 292, 00149 Rome, Italy; Department of Biology, University of Roma Tre, Viale Marconi 446, 00146 Rome, Italy, visca@uniroma3.it Ivan Visentin DiVaPRA – Plant Pathology, University of Turin, I-10095 Grugliasco, Turin, Italy Kerstin Voigt Institute of Microbiology, School of Biology and Pharmacy, University of Jena, Neugasse 25, 07743 Jena, Germany, kerstin.voigt@uni-jena.de Tapani Yli-Mattila Laboratory of Plant Physiology and Molecular Biology, Department of Biology, University of Turku, FIN-20014 Turku, Finland, tymat@ utu.fi . Part I Plant Pathological and Environmental Biological Aspects Chapter 1 Fungal Pathogens of Plants in the Homogocene George Newcombe and Frank M. Dugan Abstract As the pace of biotic homogenization has accelerated over time, the threat of novel phytopathogens has become a question of growing importance for mycologists and plant pathologists. Meanwhile, this question is but one of a whole set of related questions that invasion biologists are attempting to answer. Pathogen release is of interest to both sets of scientists because it provides a measure of the extent to which previously isolated mycobiotas have undergone cryptic homogenization, and at the same time it is the basis for a promising hypothesis to explain plant invasions. We argue that only a fraction of all first encounters between novel pathogens and evolutionarily naive plants could result in susceptible outcomes. This is analogous to the fact that only a fraction of all plant introductions result in plant invasions. 1.1 Introduction Geologists define the last 10,000 years, or our current epoch, as the Holocene (Bishop 2003). What has been described as the “Neolithic Revolution” also dates from 10,000 years ago (Wells 2007). Spurred by early developments in crop domestication in regions such as the Fertile Crescent (Wells 2007), Neolithic farmers began to move to and settle in new areas with their crops ten millennia ago (Vaughan et al. 2007). What was no doubt at first gradual and local ultimately became global. Human migrations and population expansions during the Holocene are mixing the previously isolated biotas of the world at an accelerating pace (Mooney and Cleland 2001). Organisms outside their native ranges bear many G. Newcombe Department of Forest Resources, and Center for Research on Invasive Species and Small Populations, University of Idaho, Moscow, ID 83844-1133, USA e-mail: georgen@uidaho.edu F.M. Dugan USDA-ARS, Washington State University, Pullman, WA 99163-6402, USA Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_1, # Springer-Verlag Berlin Heidelberg 2010 3 4 G. Newcombe and F.M. Dugan descriptors: non-native, nonindigenous, exotic, introduced, or alien. Non-native pathogens are additionally called “novel.” Some non-native organisms have proven invasive, and invasion biologists have begun to describe the latter part of our epoch as the Homogocene a term coined by Gordon Orians (Rosenzweig 2001a). This term invokes the global scope and increasing rate of anthropogenic, biotic homogenization that is defined as the “gradual replacement of native biotas by locally expanding non-natives” (Olden et al. 2004). The consequences of homogenization for biotic communities and ecosystem processes are the subject of a growing research effort; here we focus on fungal pathogens in the Homogocene. Unsurprisingly, most of the literature of invasion biology focuses on plants and animals (Pyšek et al. 2006), leaving mycologically oriented ecologists to wonder about fungi. Crop pathogens are of course exceptional in this regard as they are often discussed in the phytopathological literature (Rossman 2009; Stukenbrock and McDonald 2008) and lists of such pathogens that are thought to be non-native are frequently compiled (Madden 2001). The most famous historical example is arguably the pseudo-fungus (oomycete) associated with the Irish potato famine, Phytophthora infestans. Both the host and pathogen probably originated from the Andes (Gómez-Alpizar et al. 2007), and their reunion in Ireland proved disastrous both for the crop and the people who depended on it. Other examples include black Sigatoka and yellow Sigatoka of banana (causal agents Mycosphaerella musicola and M. fijiensis, respectively). In these cases, the host is from Southeast Asia, and this may also be true for these fungi that were nevertheless first documented in Fiji; both diseases now constitute global epidemics on banana (Marı́n et al. 2003). Other instances from agriculture are concisely mentioned here and there in what follows. However, this review emphasizes examples of plants from natural plant communities (i.e., nonagricultural settings). Chestnut blight and white pine blister rust (discussed below) are widely investigated, but we present many other instances much less familiar to the scientific public. We believe that a careful examination of the behavior of fungal phytopathogens in the Homogocene reveals ecologically significant patterns. These patterns should be of as much interest to invasion biologists and ecologists as they are to mycologists and plant pathologists, because fungi influence plants even while going unnoticed. Voyaging peoples have likely always brought useful and favored plants and animals with them to be deliberately introduced to new lands that they encountered. Before Captain Cook landed in Hawaii in 1778, Polynesian seafarers had already discovered these remote, volcanic islands and introduced such plants as candlenut (Aleurites moluccana), ti (Cordyline fruticosa), taro (Colocasia esculenta), sweet potato (Ipomoea batatas), sugarcane (Saccharum officinarum), and perhaps two dozen other species (Carlquist 1980). Such early plant introductions were certainly not limited to Polynesia, or to plants brought with European colonists to the New World. Plant-hunting expeditions were probably initiated as early as 3,000 years before Columbus, as hieroglyphs appear to show that the Queen of Egypt sent out collectors of exotic plants for her gardens (Baskin 2003). Mycologists know that at least some endophytic and phytopathogenic fungi must have quietly accompanied these plant introductions (Palm 1999, 2001; Palm and Rossman 2003). But what can we say today about these co-introductions and their effects on Hawaiian 1 Fungal Pathogens of Plants in the Homogocene 5 and other ecosystems around the world? Were those co-introduced endophytes so host-specific that we can be sure that they remained exclusive to their original hosts? If they jumped to other hosts, presumably related ones (Gilbert and Webb 2007), what ecological effects might they have had, and how can we distinguish them today from co-evolved interactions? These questions go well beyond Hawaii, and include all fungi globally. Maritime explorations by western Europeans that began in the early fifteenth century (Love 2006) were initially restricted in scope, but they gradually became global and much more ambitious. Scientific travelers such as Alexander von Humboldt, Charles Darwin, and Alfred Russell Wallace began in the nineteenth century to discover how life’s diversity was distributed around the world. At the same time, plant explorers of many nationalities were seeking to deliberately introduce desirable plants to their home countries (Reichard and White 2001). Crop plants had already been introduced almost everywhere that they would grow profitably; Columbus, for example, wasted no time in introducing one of the most important New World domesticates, maize, into Europe in 1493 (Rebourg et al. 2003); during the same year Columbus introduced orange trees of Asian origin to Hispaniola (Haughton 1978). Ten thousand years ago, homogenization was undoubtedly not occurring at modern rates, although it had probably been initiated on local or regional scales. So, the Homogocene cannot be said to have begun with the Holocene. Instead, 1500 would appear to be a good year to choose for the beginning of the Homogocene. This date is also in agreement with the judgment of ecological, economic, and social historians, whose titles or subtitles accordingly include phraseology suggestive of worldwide movement of goods and peoples beginning at this time, e.g., “Accumulation on a World Scale,” “Expansion of Europe,” “Modern WorldSystem,” etc. (Amin 1974; Crosby 2004; Wallerstein 1974, 1980, 1989). Five hundred and nine years ago, the “Age of Discovery” was under way, and today, deliberate plant introductions are so common that the “majority of woody invasive plants in the United States were introduced for horticultural purposes – one study found that 82% of 235 woody plant species identified as colonizing outside of cultivation had been used in landscaping” (Reichard and White 2001). The “Homogocene” may not yet be a serious term in science but it does simply and directly evoke global commingling during the last half a millennium. Plant-oriented invasion biologists in Europe already use 1500 to divide alien plants (Pyšek et al. 2004) into those that are called “archaeophytes” if introduced before 1500, and “neophytes” if introduced later. However, we hasten to make explicit what we have already hinted at: migration or purposeful import of plant materials well prior to 1500 probably made important contributions to homogenization on a regional, and sometimes even continental or oceanic, scale. In addition to the example of Polynesia above, expansion of Neolithic farming cultures such as the Arawak (from the upper Amazon and Orinoco basins to the West Indies), Bantu (from western to southern Africa), and Indo-European (from a still disputed location, perhaps Anatolia or the European steppes, but eventually throughout Europe and much of western and central Asia) moved plant materials considerable distances (Diamond and Bellwood 2003). Pronounced effects have been postulated for movement of plant pests and diseases in these distant times, e.g., the “honeymoon 6 G. Newcombe and F.M. Dugan hypothesis” of a comparatively pest- and disease-free agriculture in Neolithic Europe (Dark and Gent 2001), or the introduction of Ascochyta blight of chickpea, resulting in summer cropping systems in the Levant (Abbo et al. 2003). Literature on movement of plant pathogenic fungi from the Neolithic through classical antiquity has been summarized recently (Dugan 2008). Archaeobotanical or textbased analyses are particularly numerous for tracing the complex introduction of crops into medieval Europe (Behre 1992; Campbell 1988; Harvey 1984, 1992; Kroll 2005 ; Preston et al. 2004; Taavitsainen et al. 1998). Introduced plants are not all equal ecologically. Introduced or alien plants can become naturalized if they survive and regularly reproduce outside of cultivation (Richardson et al. 2000b). Of course, only a small fraction of introduced plants become naturalized. For example, in Florida, of approximately 25,000 non-native or alien plant species, only 900 have become naturalized (Pimentel et al. 2005). A further winnowing occurs as only a small fraction of naturalized species become invasive, with invaders defined as species that have successfully spread away from sites of introduction (Richardson et al. 2000b). These successive winnowings characterize what is called the “tens rule” (Williamson and Fitter 1996), a rule of thumb that reflects the fact that relatively few aliens become invaders (Kolar and Lodge 2001). Plant and community ecologists are keenly interested in understanding this phenomenon. In this chapter, we shall see how well concepts and definitions borrowed from invasion biology might apply to fungi, especially fungal pathogens of plants. Is there a “tens rule” for fungal pathogens, or are all alien fungal pathogens equally likely to attack evolutionarily naive plants or a host from which they had been separated? In describing the plants and pathogens that take part in “first encounters” as evolutionarily “naive” and “novel,” respectively, we are following the example of Parker and Gilbert (2004). If the “tens rule” does apply, do we have, or can we develop, hypotheses to predict which fungal pathogens will naturalize and which will become invaders? The threat of fungi as novel pathogens is a traditional topic for plant pathologists and mycologists (Rossman 2001). But, apart from the notorious example of chestnut blight, do novel pathogens generally act as “transformers” that “change the character, condition, form or nature of ecosystems over a substantial area” (Pyšek et al. 2004)? What are the roles of fungi as potential facilitators of plant invasions in the Homogocene? Recognizing and predicting invasions are the central objectives of invasion biology (Kolar and Lodge 2001). But both objectives seem to be predicated on knowing the native, geographic ranges of the organisms in question, a problematic area for mycology. 1.2 Native Ranges of Fungi In the eighteenth century, the French naturalist, Georges Buffon, had observed that different continents had different assemblages of macrobes (i.e., plants and animals) (Cox and Moore 2005). In the nineteenth century, Humboldt had 1 Fungal Pathogens of Plants in the Homogocene 7 discovered the predictability of species area relationships in that larger areas held more species (Rosenzweig 1995), but again this was known to apply only to macrobes. Spatial scaling and diversification of fungi were little studied until recently when species-area relationships of fungi were demonstrated to be similar to those of macrobes (Green et al. 2004). This was not a trivial finding because even today some microbiologists maintain the view that microbial eukaryotes have global ranges (Fenchel and Finlay 2004). The views of Beijerinck, that species of bacteria were cosmopolitan, or of Baas-Becking, that “everything is everywhere” (Fenchel and Finlay 2004), have also been challenged recently by application of the sequence-based phylogenetic species concept of fungi (Taylor et al. 2006). Mycologists are now learning that most fungi do conform to Buffon’s Law and to spatial scaling rules for macrobes. However, it does not follow that it will be easy to determine the native ranges of those fungi that do conform, for reasons that will be discussed. And then there are undoubtedly fungi that do not conform. For example, some saprophytic hyphomycetes, such as common Cladosporium species, are associated with a very broad range of substrata. Such species do indeed seem to have cosmopolitan distributions as evidenced by their incorporation over long time periods into Arctic ice, alpine glaciers, and permafrost throughout the Northern Hemisphere (summarized in Dugan 2008). Macrobiologists may be surprised to learn that the native ranges of fungi are largely unknown. Yet, how could it be otherwise? Today, 83% and 90% of vascular plants and vertebrates, respectively, are known to macrobiologists, whereas, at best, from 7% to perhaps 20% of the fungi are presently described (Cox and Moore 2005; Hawksworth 2001; Rossman 2009). Now when a new species of fungus is described, its current, geographic pattern of occurrence might suggest an original native range. Unfortunately, that pattern could also be the product of homogenization since 1500. In contrast, the native ranges of macrobes are largely known, not only because the species are largely known, but because their ranges were documented early in the Homogocene before homogenization had had large effects. Disputes do exist, but they appear minor in scope to a mycologist. For instance, Gayther Plummer proposed that the most mysterious of native trees of North America, Franklinia alatamaha, or the Franklin tree, was actually introduced from Asia a few decades before the Bartrams discovered a small grove in 1765 (Rowland 2006). Most botanists, however, disagree with Plummer (USDA n.d.). The native ranges of annual brome-grasses have more recently presented more serious challenges to botanists (Smith 1986), and other examples exist of course, but botanists have a set of criteria for dealing with problematic taxa: paleobotanical evidence of native status, records of their presence in their current range by early botanists, and current presence in natural habitats (Pyšek et al. 2004). Mycologists face the unknown species problem, and the problem of the near total lack of knowledge of pre-Homogocene distributions. Mycologists were not on board the ships of the explorers and palaeomycology can hardly arbitrate disputes, as it is “in its infancy” (Stubblefield and Taylor 1988); others have even argued that the fungi lack “any significant fossil record” (Cain 1972). Even today fungi 8 G. Newcombe and F.M. Dugan are more intensively studied in managed habitats, as pathogens of agicultural crops, than in natural habitats where fungi provide ecosystem services on a massive scale. Last, but not least for a book on the molecular identification of fungi, “molecular diagnostic tools are only as good as the systematic underpinnings upon which these tools are based” (Rossman 2009), and upheavals in fungal systematics are common today. What can we make of a new species such as Cladosporium subtilissimum that was described recently from material in Slovenia and the northwestern United States (Schubert et al. 2007)? It could have been cosmopolitan prior to the Homogocene, but can we rule out the role of homogenization in producing its current distribution? Climate change can of course also cause range shifts (Parmesan 2006), but homogenization is more likely to be the source of the error that we are concerned with here: calling an invaded or naturalized range a native range or part of a native range. How can this error be avoided? And to what extent have previously isolated mycobiotas already been cryptically homogenized? 1.3 Pathogen Release A roundabout but fruitful way to approach the latter question is through the pathogen release hypothesis, according to which alien plants are less regulated by pathogens than native plants (Keane and Crawley 2002). But first, if all fungi were everywhere, as Beijerinck, Baas-Becking, Fenchel, and Finlay have asserted is the case for other microbes (Fenchel and Finlay 2004), there would be no pathogen release for plants from plant pathogenic fungi. Plants would have the same set of fungal pathogens in both their native and invaded ranges. The environment (i.e., the host plant) would select. Is pathogen release a real phenomenon? Using the USDA Fungus–Host Distributions database of the Systematic Mycology and Microbiology Laboratory, Mitchell and Power showed that for “473 plant species naturalized to the United States from Europe” there were 84% fewer rust, smut, and powdery mildew species infecting plants in their naturalized ranges than in their native ranges (Mitchell and Power 2003). The SMML database is by far the most extensive of its kind with “reports of fungi on plant hosts throughout the world that includes over 94,000 fungal species” (Rossman 2009). One could also cite specific examples of fungi that have been deliberately introduced for classical biocontrol of weedy plants that more directly confirm the pathogen release hypothesis for phytopathogenic fungi (Cullen et al. 1973), but Mitchell and Power’s paper was the first study to show the generality of this phenomenon. Another way of phrasing this is that if phytopathogenic fungi were everywhere, then pathogen introductions would not be a threat. To prove that this belief represents a completely false sense of security, one has to look no further than the chestnut blight fungus that transformed an ecosystem (Cox 1999; Liebhold et al. 1995; Rizzo and Garbelotto 2003). 1 Fungal Pathogens of Plants in the Homogocene 9 Although plants may at first leave their fungal enemies behind when introduced outside their native ranges, in keeping with pathogen release, one can imagine that inadvertent introductions of those same enemies would slowly counter the pathogen release effect over time. These “pathogen reunions” do occur, and they are perhaps the best measure that we have of the rate of introduction of fungi (i.e., fungal homogenization) around the world. For example, when rust occurred on Centaurea diffusa, an invasive plant of Eurasian origin, for the first time in North America in 1989 (Mortensen et al. 1989) or for the first time in the United States in 1992 (Dugan and Carris 1992; Palm et al. 1992), these pathogen reunions ended more than 80 years of release from rust dating from 1907, the year that C. diffusa itself was introduced into North America (Maddox 1982). This rust, Puccinia jaceae var. diffusae of Eurasian origins (Savile 1970b), is easily distinguished from the only rust fungus, Puccinia irrequiseta, that occurs on the only North American species of Centaurea, C. americana (Savile 1970a). In the case of tansy, or Tanacetum vulgare, plants were introduced and cultivated by English colonists in North America for culinary and medicinal purposes (Haughton 1978). Rust, common in its native range, was absent from this introduced range. Some 400 years after the introduction of tansy, Puccinia tanaceti was finally reunited with its host for the first time in the North American range of tansy (Newcombe 2003b). Cochliobolus carbonum provides an example of serial pathogen reunions in that it must have followed the introduction of its host, Zea mays, around the world to the point where the fungus itself is now cosmopolitan. Some of the reunions were relatively recent. For instance, C. carbonum only reached Great Britain in 1972 (Jones and Baker 2007), which is presumably long after Z. mays was introduced there, as the plant was introduced into Europe in 1493 by Columbus (Rebourg et al. 2003). C. carbonum had reached Australia 6 years before it arrived in the U.K. (Farr et al. n.d.). Soybean rust, caused by Phakopsora pachyrhizi and P. meibomiae, also took considerable time to be reunited with its agriculturally important host around the world (Rossman 2009). It is not clear how lengthy periods of pathogen release might potentially be as many pathogen reunions have yet to occur. Morus alba, the white mulberry, was deliberately introduced from China in an attempt to establish a silk industry in the U.S. more than 400 years ago (Duncan and Duncan 1988). The tree has naturalized, and even become locally invasive in the U.S., but its rust fungi (i.e., species of Cerotelium, Peridiopsora, Phakopsora, and Kuehneola) have remained in the native range of their host (Farr et al. n.d.). Powdery mildews, on the other hand, have reunited with introduced populations of white mulberry in western Europe and Central America, although not yet in North America (Farr et al. n.d.). Similarly, St.-John’s-wort, Hypericum perforatum, was introduced into North America by Rosicrucian pilgrims in 1696 (Haughton 1978), and it has since become weedy and invasive across the entire continent (USDA n.d.). But more than 400 years later Melampsora hypericorum has yet to be reunited with H. perforatum, 10 G. Newcombe and F.M. Dugan as this rust fungus has only been recorded in St.-John’s-wort’s native range in Europe (Farr et al. n.d.). On the other hand, it was close to 2,000 years ago that the Romans introduced H. perforatum to the U.K. (Haughton 1978), where reunion with M. hypericorum eventually did occur sometime before 1913 (Grove 1913 ). The “honeymoon hypothesis” of Dark and Gent (2001), mentioned above, posited that some reunions of grave consequence to European agriculture were postponed for centuries, but these reunions eventually took place as long distance movement of seeds became more routine in the late Iron Age and Roman times. Plants native to North America were also introduced to Europe where some remain in a state of at least partial pathogen release. Helianthus tuberosus, the inappropriately named “Jerusalem artichoke,” was brought to Europe from North America in the early 1600s (Hedrick 1950). Although Puccinia helianthi was then reunited with its host in western Europe nearly 400 years later, other rust fungi remain restricted to the native range (e.g., Coleosporium helianthi) (Farr et al. n.d.). A similar pattern of pathogen release is known for Helianthus annuus, the cultivated sunflower, also a native of North America that became widely cultivated around the world. It is important to note that pathogen reunions may be confused with infection by morphologically similar fungi that are native to the naturalized range of an introduced plant. For instance, Populus nigra is Eurasian, but it has been widely planted in North America as cv. ‘Italica,’ the columnar Lombardy poplar. Venturia infection of P. nigra in North America could potentially represent pathogen reunion by a Eurasian Venturia, or host switching by a Venturia that is native to North American Populus. As it turns out, Venturia populina, a Eurasian fungus, was determined to be causing leaf and shoot blight of P. nigra (Newcombe 2003a), so this was a case of pathogen reunion. Venturia inopina, occurring on a North American species of Populus, P. trichocarpa, is morphologically similar to V. populina, but it has not switched to P. nigra. It was the specificities of these two species of Venturia for their respective hosts, expressed in a common environment, that actually led to discernment of subtle, but consistent, differences in morphology and in ITS sequences. In retrospect, Venturia blight of P. nigra in North America could easily have been misinterpreted as host switching, or simply as the product of a fungus with a broader host range and larger geographic range than either of these species of Venturia actually has. In general, most of the introductions of alien or so-called “invasive” plant pathogens appear to be pathogen reunions. In a recent study of 1970–2004, among the introductions of non-native plant pathogens into the U.K., 85% were reunions on plants that were themselves introduced. Only 15% were first reports of pathogens on native, wild plants of the U.K., and not all of these were necessarily reports of alien pathogens (Jones and Baker 2007); some could have been native pathogens that had been overlooked because their hosts lacked economic importance. Pathogen reunions may be the best measure that we have of the rate of introduction of fungi, but what do they tell us about the native ranges of fungi, the primary question of this section? 1 Fungal Pathogens of Plants in the Homogocene 1.4 11 Inferring Native Ranges of Fungi from Pathogen Release It is tempting to think that fungi with restricted host ranges must be native where their hosts are native. Camellia, a genus of some 200 species, is endemic in eastern Asia, with its center of diversity in southern China (Ta and Bartholomew 1984). As Ciborinia camelliae is restricted to Camellia, its discovery on C. japonica in Great Britain in 1999 (Jones and Baker 2007) should ultimately be traced back to a native range in eastern Asia even though other parts of the world may have been stepping stones. If species of Camellia had never been introduced outside eastern Asia, the inference of sympatry for its host-restricted fungi would be unequivocal. Ornamental species of Deutzia provide such an example in that they are also endemic to Asia where seven taxa of rust fungi commonly infect them (Farr et al. n.d.). Unlike C. camelliae, records of rust on Deutzia outside the native range are absent even though D. scabra was introduced to the U.S. as early as 1822 (Rehder 1940). Searches of the SMML Fungus–Host Distribution Database (Farr et al. n.d.) suggest restricted host and geographic ranges of fungi too numerous to comprehensively list here, but it is instructive to provide examples. Three species of Pucciniastrum occur only on Asian species of Acer, maple, and only in Asia. Rust occurs on the English oak, Quercus robur, in its native range but not in its introduced range in North America, even though Uredinales is well represented on North American Quercus. Amelanchier alnifolia supports 17 rust taxa in its native range in North America, but none in Europe where it has naturalized (Zerbe and Wirth 2006). Presence in the host’s native range and absence in its naturalized range allow for strong inference of the native range of a fungus (Table 1.1); presence in both of the host’s ranges is problematic only in the absence of historical records of absence of the fungus in one of them. It is even more tempting to think that the combination of Fahrenholz’s rule and knowledge of the native ranges of plants can be used to further strengthen inferences of native ranges of fungi. Fahrenholz’s rule postulates that “parasites and their hosts speciate in synchrony” (Hafner and Nadler 1988). If host switching were not an issue, then native ranges of hosts should also be native ranges of their parasites. However, host switching is an issue (Jackson 2004). Host switching is best exemplified by absence in the host’s native range and presence in its naturalized range. This requires some explanation, aided by the example of Eucalyptus rust (Grgurinovic et al. 2006). Puccinia psidii causes Eucalyptus rust but the first reports of this disease were not from the native range of species of Eucalyptus in Australia. Instead, this rust fungus was first reported on plantations of eucalypts grown in Brazil. Evidently, P. psidii had switched, or jumped, from species of Myrtaceae native to South America to introduced species of Eucalyptus that also belongs to Myrtaceae. Host switching can also be inferred from the early years of agriculture, e.g., for formae speciales of Blumeria graminis. Strict coevolution was apparently absent between this 12 G. Newcombe and F.M. Dugan Table 1.1 Three categories of first encounters between evolutionarily naive plants and novel pathogens that depend on two factors: (1) which party to the encounter is alien, and (2) whether opportunities for encounters will be prolonged or brief. The category of the encounters in turn affects how susceptible and resistant outcomes of first encounters contribute, or not, to biotic resistance Time Evolutionarily naive parties to first encounter Biotic resistance (BR) of native biotic community versus alien plants or alien pathogens Examples of encounters Naive Plant Novel Pathogen Expected outcome of Category Opportunities for encounters (discussed in text) encounters, if (prolonged by pathogen contributing to BR reunion) 1 Extended, as naturalized, alien Alien Native Susceptible (S) R (Prunus serotina resistant to Uredinales in Europe). S plants remain exposed to (individuals of Pinus pathogens of native plants (no) sylvestris susceptible to Endocronartium harknessii in North America) 2 Extended, when alien Native Alien Resistant (R) R (6 of 7 taxa of Malus resistant to Podosphaera leucotricha pathogens are reunited with alien, naturalized plants in N. America). S (Malus (yes) angustifolia, the seventh taxon, susceptible to P. leucotricha in N. America) 3 Short (no) Native Alien Resistant (R) R (North American pines with Cr genes for resistance to Cronartium ribicola, an Asian fungus). S (individuals lacking these genes, in the same North American pines) 1 Fungal Pathogens of Plants in the Homogocene 13 powdery mildew and its grass family hosts in western Asia (Wyand and Brown 2003). Likewise, the barley scald pathogen, Rhynchosporium secalis, apparently evolved on other hosts outside the center of diversity for barley (Zaffarano et al. 2006). 1.5 Genetic Criteria for Native Range Host range is sometimes not specific enough to even suggest a particular native range for a fungus. For example, Venturia inaequalis, the apple scab pathogen, affects all species of Malus, some of which are native to North America although most are Eurasian. Records of occurrence of V. inaequalis might be misleading in that the apple, first domesticated in central Asia, was introduced by early explorers everywhere that it would grow (Hedrick 1950). It is common in such cases to hypothesize that genetic variation will be greatest in the native range. Using this criterion, Gladieux et al. showed that V. inaequalis is likely native to the same area in Asia in which apple itself was domesticated (Gladieux et al. 2008). Similarly, an Asian origin of the dry rot fungus, Serpula lacrymans, has been inferred from a study of its genetic variation (Kauserud et al. 2007). Genetic variation also places the amphibian chytrid fungus, Batrachochytrium dendrobatidis, in a native range in South Africa (Weldon et al. 2004) from where it has spread to cause a pandemic. Interestingly, the highest genotypic diversity for the human dermatophyte, Trichophyton rubrum, is in Africa, where Homo sapiens itself evolved (Gräser et al. 2007). Genetic diversity is not likely to be by itself an infallible criterion of native range, however. Plants with outcrossing mating systems are frequently as genetically diverse in their naturalized or invaded ranges as they are in their native ranges (Novak and Mack 2005). Ambrosia artemisiifolia, a North American plant, maintains high genetic diversity in its invaded range in France (Genton et al. 2005). Studies of Bromus tectorum have shown that even selfing plants may be as diverse in the invaded range, in North America, as in the native, Eurasian range, if the invasion involved multiple introductions (Novak and Mack 2005). A recent summary of 20 analyzes of genetic diversity in invasive plant populations showed that estimates of total genetic diversity vary from “none detected” to “high” (Ward et al. 2008). Similar caveats may apply to sole use of a genetic criterion for determination of native ranges of fungi; genetic diversity of pine-associated Sphaeropsis sapinea is high in South Africa where the fungus must have been introduced from the northern hemisphere (Smith et al. 2000). Conversely, North American populations of Entoleuca mammata are genetically more variable than introduced populations in Europe (Kasanen et al. 2004). The oak wilt fungus, Ceratocystis fagacearum, is only known to occur in the middle and eastern United States, but its genetic homogeneity has led some researchers to hypothesize an exotic origin; the oak populations of Mexico or Central America have been suggested (Juzwik et al. 2008). An alternative hypothesis to explain low diversity is “the local genesis of a new and reproductively isolated strain or species” (Zambino and Harrington 2005). 14 G. Newcombe and F.M. Dugan However, low genetic variation can also characterize ancient and relictual species. For example, the Wollemi pine, Wollemia nobilis, is the only extant member of its genus, surviving only as a single, small population in a canyon in Australia (Peakall et al. 2003). No genetic variation whatsoever has been detected in W. nobilis. Genetic drift can reduce genetic variation in small and isolated populations of plants (Ouborg et al. 2006), and of organisms generally, but this explanation for genetic homogeneity of fungal species has not been widely considered. 1.6 Inferring Native Ranges of Pathogenic Fungi from Resistance Given historical examples of extreme susceptibility of plants to novel pathogens (e.g., chestnut blight, and white pine blister rust), and the resistance of related plants elsewhere, it is tempting to think that host resistance can indicate native range of a pathogen. Many plant pathologists have followed this line of thinking. Their evolutionary explanation is that of selection for resistance in the presumptive native range, and the absence of such selection elsewhere. The host in the presumptive native range is thought to be evolutionarily adapted, whereas the host in the invaded range is said to be evolutionarily naive with respect to the novel pathogen. The Asian species of Castanea and Pinus were certainly more resistant to the chestnut blight and white pine blister rust fungi, respectively, than were North American species of Castanea and Pinus. With these examples as paradigm, researchers have tried to infer the native range or origin of many other pathogenic fungi that were unknown or poorly known prior to an epidemic. The difficulties in doing so are twofold. First, the pathogen has to be present in the native range of a mostly resistant, putative host. Secondly, the resistance of the latter has to be adaptive and cannot be complete; the pathogen has to be able to survive and reproduce, so there must either be some susceptible individuals of an otherwise resistant species, or the adapted host could be tolerant (Roy and Kirchner 2000). Even then, it may prove surprisingly difficult to distinguish adaptive resistance, which is associated with ongoing selection, from exapted resistance (Newcombe 1998). Exaptations are characters that in a new evolutionary context can have selective value even if they resulted from selection for something else (Gould and Vrba 1982). For example, the resistance of Populus maximowiczii, a poplar native to eastern Asia, to species of Venturia and Taphrina on Canada’s Vancouver Island could only have been construed as adaptive if there had been evidence that those fungi were native to eastern Asia. Needless to say, that would not have been a parsimonious interpretation (Newcombe 2005). The evolution of plant-pollinator and plant-herbivore interactions may be linked through exaptations (Armbruster 1997). However, the linkages of exaptations for resistance to pathogens are unknown. One can only speculate that the function of genes in a poplar species in eastern Asia for resistance to Vancouver Island fungi 1 Fungal Pathogens of Plants in the Homogocene 15 would presumably be related to defense of some kind. Adaptive resistance must also be distinguished from nonhost resistance that is predicted, but not explained, by phylogenetic signal (Gilbert and Webb 2007; Newcombe 2005). Unfortunately, little attention has been paid to these distinctions when inferring native ranges of plant pathogens. In discussing root rot of Port Orford cedar caused by Phytophthora lateralis, the assumed relationship between resistance and origin of a pathogen was stated in this way: “P. lateralis was described in 1942. It is suspected to be of Eurasian origin because Asiatic species of Chamaecyparis resist it; resistance may have arisen through coevolution with the pathogen (Sinclair et al. 1987).” Eighteen years later however, Sinclair adopted a much more cautious position: “P. lateralis, origin unknown” (Sinclair and Lyon 2005). This caution in inferring native ranges of pathogens from the geographic distribution of resistance is warranted as the examples which follow will hopefully make clear. In the instance just discussed, the nature of resistance (i.e., adaptive versus exapted/ nonhost) of Asian Chamaecyparis to P. lateralis remains unclear. Dogwood anthracnose, for example, is also thought to be caused by a highprofile, alien pathogen, Discula destructiva. Discula destructiva has been regarded as an alien pathogen in North America (Redlin 1991), that has more recently appeared in the U.K. (Jones and Baker 2007), and in western Europe (Holdenrieder and Sieber 2007). Its appearances in the U.K. and Europe have been on Cornus florida that is native to North America. Susceptibility associated with severe damage and mortality in nature is only seen in the North American species, C. florida and C. nuttallii (Sinclair and Lyon 2005). Inferring the native range of D. destructiva from these observations, one would conclude that its origins were Eurasian. But, actually, resistance is not a helpful criterion because it characterizes at least some species of Cornus native to each one of the three continents where D. destructiva could be native: North America, Europe, and Asia (Holdenrieder and Sieber 2007; Sinclair and Lyon 2005). Although the native range of D. destructiva has been hypothesized on the basis of host resistance to specifically coincide with that of Cornus kousa in eastern Asia (Redlin 1991), the fungus has never been reported on C. kousa in its native range. Fourteen fungi have been recorded on C. kousa in eastern Asia (Farr et al. n.d.), but D. destructiva is not one of them. Interestingly, C. florida, grown as an introduced ornamental in Japan, has proven susceptible there to endemic Pucciniastrum corni, a number of taxa of Erysiphales, and a few other fungi, but D. destructiva has not been recorded on it. This evidence of absence in Japan is not definitive but it does suggest that resistance-based inferences of origin can be quite misleading. Butternut canker provides another example. Sirococcus clavigignenti-juglandacearum is thought to be an alien pathogen in North America that might have been introduced on seed of Asian species of Juglans such as J. ailantifolia (Ostry and Moore 2007). But the only record of S. clavigignenti-juglandacearum on J. ailantifolia is from the U.S. (Ostry 1997), and records of this fungus from Asia are lacking on any host (Farr et al. n.d.). Nine fungal taxa have been recorded on J. ailantifolia in Asia, so the fungi of the putative, adaptive host are not completely unresearched (Farr et al. n.d.). 16 G. Newcombe and F.M. Dugan Records of eastern filbert blight, caused by Anisogramma anomala, are also restricted to North America (Farr et al. n.d.). But possibly because there were records of this disease from the late 1800s, A. anomala is not thought to be alien to North America (Sinclair and Lyon 2005). In fact, the native host of A. anomala is hypothesized to be Corylus americana (Coyne et al. 1998). But if resistance patterns were the sole criterion for native range, the resistance of some species of Corylus native to each of North America, Europe, and Asia would again be problematic (Coyne et al. 1998; Sinclair and Lyon 2005). Furthermore, if A. anomala is not present in Europe and Asia, and is truly native to North America, then the resistance of Eurasian species of Corylus must be of the exapted or nonhost variety rather than adaptive. As some individuals of the European hazelnut, C. avellana, are susceptible, nonhost resistance would seem to be ruled out. So, the dominant, “Gasaway” gene that is inherited from resistant individuals of the European hazelnut (Coyne et al. 1998) would be interpreted here as exapted. Seiridium cankers of cypress are caused by three species of Seiridium. Relatively resistant species of Calocedrus, Chamaecyparis, Cupressus, Juniperus, Taxodium, Thuja, and Thujopsis are native to both Eurasia and North America (Sinclair and Lyon 2005). Once again, the lack of any discrete, geographic source of resistance known to be adaptive would thwart any attempt to pin any of the three species of Seiridium to any particular native range, at least using the sole criterion of resistance. Fusiform rust, caused by Cronartium quercuum f.sp. fusiforme, is also instructive. The fusiform rust fungus is thought by many to be native to the region where it is currently most damaging: the southeastern U.S. However, after testing 45 species of Pinus for susceptibility/resistance to C. quercuum f.sp. fusiforme, an origin of C. quercuum f.sp. fusiforme in Central America was hypothesized (Tainter and Anderson 1993). This hypothesis was congruent with the relatively strong resistance of species of Pinus from Central America. However, equally strong resistance of Asian and Mediterranean species was evident in this study. To be consistent in applying the criterion of resistance, the authors would then have had to propose a native range for C. quercuum f.sp. fusiforme involving widely scattered, disjunct populations in Asia, Europe, and Central America, quite unlike that of any species of Pinus that hosts the fusiform rust fungus. The oak wilt fungus, Ceratocystis fagacearum, of the middle and eastern United States, has already been mentioned. Its genetic homogeneity has led some researchers to hypothesize an exotic origin (Juzwik et al. 2008). But, it is impossible to infer the native range of C. fagacearum from host resistance alone for no other reason than that this subject remains seriously understudied in some 530 species of Quercus (Mabberley 2008). So, in addition to the challenging need to distinguish adaptive and exapted/nonhost resistance, undersampling issues can be formidable. One might imagine that difficulties in determining native ranges of fungi from resistance are only encountered when alien pathogens are obscure and of little importance. But not only are the examples just cited important, but even fungi as important as the Ophiostoma species that have caused global pandemics of Dutch elm disease in the past century, remain of uncertain, geographic origin (Brasier and 1 Fungal Pathogens of Plants in the Homogocene 17 Buck 2001). Surveys in China (Brasier 1990) and in the Himalayas (Brasier and Mehrotra 1995) were undertaken because resistant species of Ulmus are naturally distributed there. However, O. ulmi and O. novo-ulmi.were not found in either surveyed region. Absence of the relevant pathogens implies that the resistance of Asian elms is exapted or nonhost, rather than adaptive. It should be noted that Ophiostoma species related to the Dutch elm fungus are one possible, and greatly debated, cause of the great mortality of elms in Europe during the Neolithic (summarily reviewed in Dugan 2008). The above examples should not only encourage caution in inferring native ranges of fungi from resistance. They should also provoke questions about the threat of novel pathogens of plants. Even if adaptive resistance is obviously, by definition, absent in naive plants encountering novel pathogens, could not exapted or nonhost resistance protect them, and if so, at what frequency? 1.7 First Encounters Between Evolutionarily Naive Plants and Novel Pathogens As previously mentioned, predicting outcomes of invasions is one of the central objectives of invasion biology (Kolar and Lodge 2001). With plants and animals, nonrelational hypotheses focus either on the relative invasiveness of potential invaders or the relative invasibility of potentially invaded communities (Heger and Trepl 2003). Relational hypotheses consider both. In all cases, however, it is considered essential to know the original, pre-Homogocene, geographic ranges of the organisms in question to develop and test hypotheses. To test the pathogen release hypothesis (Keane and Crawley 2002), for example, one needs to compare the enemies of a particular plant or animal in its native and invaded ranges. Novel weapons might aid a plant invader but only when wielded against species that are evolutionarily naive in the sense that they have never faced the weapons in question (Callaway and Ridenour 2004). Even Darwin’s hypothesis, the first hypothesis of invasiveness (Rejmánek 1996), that invasive species are more likely from alien genera than from genera found in both ranges is, of course, predicated on knowing what those ranges are (Darwin 1859). For fungal pathogens involved in pathogen reunions, prediction of outcomes is straightforward. Plant species “ABC,” known to be susceptible in its native range to pathogen “abc,” is likely to be susceptible everywhere else to “abc,” provided that the environment is conducive to infection and disease expression. We might assume that the native range of “abc” is the same as that of “ABC,” but all we really need to know for predictive purposes is that the latter is susceptible to the former somewhere else. But, pathogen reunions are not first encounters. Outcomes of true, first encounters between naive plants and novel pathogens appear to be much more challenging to predict. Thus far, we have been operating under the assumption that for predictive purposes the native ranges of phytopathogenic 18 G. Newcombe and F.M. Dugan fungi must be known. However, the status of a first encounter may be ascertained merely by knowing that the two parties differ in their native ranges. In some cases, we may know their native ranges: the host switching that, for instance, occurred when naive Eucalyptus was introduced to South America and novel Puccinia psidii switched to Eucalyptus from native Myrtaceae. But, in the case of Dutch elm disease or dogwood anthracnose, as discussed above, neither the native ranges of the novel pathogens nor the identities of their adaptive hosts have ever been determined. Nevertheless, it is clear that these diseases do represent first encounters that must involve host switching. Such encounters also represent the fungal component of what invasion biologists call “biotic resistance” (Parker and Gilbert 2004). Any and all organisms in a native community can provide biotic resistance to repel invaders. When pathogenic fungi switch from native plants to naive, alien plants, as in rust of Eucalyptus, they contribute to biotic resistance. This scenario represents one of three categories of first encounters between naive plants and novel pathogens that we emphasize here (Table 1.1): (1) alien plants versus native pathogens; (2) native plants versus alien pathogens involved in pathogen reunions, i.e., the majority of alien pathogens, which exist in their new, non-native environments on alien but naturalized hosts (Jones and Baker 2007); (3) native plants versus alien pathogens not involved in pathogen reunions. The first two categories allow pathogens considerable lag periods during which host switching may occur from their adaptive hosts. It is important to recall that alien organisms frequently become invasive only after considerable lag periods (Mack et al. 2000). The third category is distinct because it involves alien pathogens that must switch to the naive host immediately upon introduction because their adaptive host is absent. The null hypothesis might be argued that these distinctions of three categories are unnecessary because evolutionarily naive plants will always be decimated by alien pathogens, particularly if the plants are closely related to the host of the alien pathogen in question. The examples of white pine blister rust and chestnut blight surely suggest as much. In the absence of selection for resistance, is not genetic susceptibility to pathogens of exotic congeners inevitable and complete? The short answer is no. Much of what we have already discussed implies this. The easiest way to expand that answer is to further discuss resistance to pathogens that were clearly alien, starting with the white pine blister rust fungus, Cronartium ribicola. The latter was a “Category 3” alien pathogen in North America (Table 1.1), but with an asterisk; it was introduced without an adaptive host but on a naive host, Pinus strobus, to which it had already switched outside North America (Kinloch 2003). Cronartium ribicola is considered native to eastern Asia where its adaptive hosts are thought to include Pinus sibirica, P. armandii, P. koraiensis, P. wallichiana, and P. pumila (Kinloch 2003; Kinloch and Dupper 2002; Sinclair and Lyon 2005). The host range of C. ribicola spans the species of Pinus belonging to subgenus Strobus that includes sections Quinquefoliae and Parrya (Gernandt et al. 2005). These sections are especially speciose in North America and Asia, and less so in Europe where P. cembra and P. peuce are native. When the latter two European species, seven Asian species, and eight North American species were tested for 1 Fungal Pathogens of Plants in the Homogocene 19 blister rust resistance in Europe, only P. cembra and the Asian species, P. armandii and P. pumila were completely resistant (Stephan 2001). Among the North American species, P. aristata of section Parrya was more resistant than the seven species representing section Quinquefoliae: P. strobiformis, P. balfouriana, P. lambertiana, P. albicaulis, P. flexilis, P. monticola, and P. strobus. When blister rust resistance was tested in North America, P. strobiformis was more resistant than other species of North American origin (Sniezko et al. 2008). The point of emphasis here is the fact that none of the North American species encountering C. ribicola for the first time were completely susceptible because resistant individuals have been found in each (Stephan 2001; Sniezko et al. 2008). In fact, four species (i.e., P. strobiformis, P. monticola, P. flexilis, and P. lambertiana) have been shown in separate studies to possess major genes for resistance, albeit at low frequencies (Kinloch 1992; Kinloch and Dupper 2002; Kinloch et al. 1999, 2003). Although phenotypic evidence for these Cr genes was not detected in whitebark pine (P. albicaulis), Mexican white pine (P. ayacahuite), foxtail pine (P. balfouriana), and Great Basin bristlecone pine (P. longaeva), all of these species might possess such genes at low frequencies that would simply require additional sampling for their discovery (Kinloch and Dupper 2002). The authors conclude that although “blister rust traditionally is considered an exotic disease in North America, these results, typical of classic gene-for-gene interactions, suggest that genetic memory of similar encounters in past epochs has been retained in this pathosystem” (Kinloch and Dupper 2002). Before considering examples of exapted resistance other than the Cr genes, the implications of gene-for-gene interactions require explanation. Disease resistance in plants is often tackled by using some conceptual dichotomy. Van Der Plank famously distinguished between vertical and horizontal resistance, for example (Van Der Plank 1975). Gene-for-gene interactions characterize Van Der Plank’s vertical resistance (Briggs and Johal 1994; Flor 1971; Thompson and Burdon 1992). Flor (1971) developed the gene-for-gene theory by performing correlated studies of the inheritance of both host resistance and pathogen virulence using cultivated flax and flax rust. Flor is typically quoted for defining these interactions in this way: “for each gene that conditions reaction in the host there is a corresponding gene that conditions pathogenicity in the pathogen.” Gene-for-gene interactions can also be inferred, somewhat less rigorously (Thompson and Burdon 1992), by proving that there are a number of distinct, major genes for resistance that allow pathogen isolates to be differentially distinguished as pathotypes. By this definition, gene-for-gene interactions do characterize white pine blister rust (Kinloch and Dupper 2002) and also poplar leaf rust that we shall discuss next as it also involves the sudden appearance of “genetic memory” in first encounters (Newcombe et al. 2001). Recall that gene-for-gene interactions are thought to be the product of continuous coevolution (Person 1959, 1967). But both the poplar leaf rust pathosystem of the Pacific Northwestern region and the blister rust pathosystems of the white pines of North America appear to be the product of recent pathogen introductions, exapted resistance genes, and recent, adaptive changes in the pathogen populations. 20 G. Newcombe and F.M. Dugan Attempts to explain the Cr genes in terms of selection have been made; the highest frequencies of Cr1 and Cr2 are in the American Southwest near overlaps with pinyon pines of section Parrya and pinyon blister rust caused by an American native rust fungus, Cronartium occidentale (Kinloch and Dupper 2002). However, to positively demonstrate that C. occidentale was the selective agent that explains evolutionary retention of Cr genes, one would have to show that Cr genes protect species in section Quinquefoliae against C. occidentale (Kinloch and Dupper 2002). Otherwise, the retention of Cr genes in North American white pines appears paradoxical given the absence of selection (Kinloch and Dupper 2002). In the case of poplar leaf rust also, resistance genes were revealed by pathogen introductions, raising again the question of their retention in the absence of selection. Complex pathogenic variation indicative of gene-for-gene interactions also appeared very quickly in this system, once the pathogen population had undergone hybridization to match that of its hybrid host (Newcombe et al. 2001). Some background is needed to explain current gene-for-gene interactions in poplar leaf rust in the Pacific Northwest of North America. Populus trichocarpa, the western black cottonwood, is native to the region, along with its coevolutionary rust, Melampsora occidentalis. When P. trichocarpa (T) from section Tacamahaca is crossed with P. deltoides (D), the eastern cottonwood, from section Aigeiros, fastgrowing F1 hybrid clones can be selected. These TxD hybrids have been the mainstay of commercial poplar plantations in the region for nearly three decades. Initially TxD hybrids were rust-free. The resistance of P. deltoides to M. occidentalis (Newcombe et al. 2000), was transmitted to all TxD F1 hybrids indicating that these P. deltoides parents are dominant homozygotes in this respect. In 1991, Melampsora medusae, the coevolutionary rust of P. deltoides, was found in the region. It quickly became apparent that some TxD F1 hybrids were susceptible to M. medusae. Analysis of the inheritance of resistance to M. medusae in a TxD hybrid poplar pedigree demonstrated that the Mmd1 gene for resistance was inherited from the P. trichocarpa parent (Newcombe et al. 1996). It is important to note that this gene had gone unnoticed in previous studies of the resistance of P. trichocarpa to M. occidentalis. The gene-for-gene explanation for the detection of Mmd1 with M. medusae is that the latter evidently possesses the matching avirulence allele, unlike the coevolutionary rust, M. occidentalis. Both P. trichocarpa and M. medusae could even be fixed for this gene-for-gene pair as their interaction phenotype was always resistant, although testing was limited to nine individuals of the former and four isolates of the latter from the southeastern U.S. (Newcombe et al. 2000). Resistance to M. medusae was observed to segregate in the TxD F1 because the P. trichocarpa parent was heterozygous at Mmd1 (Newcombe et al. 1996). Until 1995, there was no pathogenic variation in the rust population that simply consisted of M. medusae. F1 clones either possessed the dominant Mmd1 allele for resistance, or not. Emergence in the mid-1990s of the hybrid of M. medusae and M. occidentalis, M. x columbiana, changed this situation. Previously resistant TxD F1 clones became susceptible. It rapidly became apparent that there was abundant pathogenic variation in the new hybrid population of M.  columbiana 1 Fungal Pathogens of Plants in the Homogocene 21 (Newcombe et al. 2001). Just as M. medusae had allowed the Mmd1 gene to be detected, new pathotypes of M.  columbiana were the means by which three new genes for resistance, Mxc1, Mxc2, and Mxc3 were discovered. A new gene-for-gene pathosystem had appeared with exapted resistance genes and matching exapted avirulence genes. Resistance genes appear to be quite common in Populus, perhaps totaling in the hundreds (Tuskan et al. 2006). Hybridization has been hypothesized to be a stimulus for the evolution of invasiveness in plants (Ellstrand and Schierenbeck 2000), and it may be so also for pathogenic fungi (Brasier 2000). Hybridization of both host and parasite that merged two separate pathosystems appears to account for the emergence of this gene-for-gene system. Reciprocal hybridization of the kind discussed here could have had an evolutionary history of repeated occurrence, as there is evidence of ancient hybridization between Populus sections Tacamahaca and Aigeiros, at least since the Miocene (Eckenwalder 1984). The genes for resistance and avirulence that now appear exapted could have been selected episodically in the past in recurring, hybrid zones. Populus trichocarpa and P. deltoides do currently hybridize naturally in parts of western North America (Eckenwalder 1996). The ancient introgression of Pinus banksiana into Pinus contorta in western North America (Critchfield 1985), has also left a signal in terms of resistance genes, that is still evident today (Wu et al. 1996). But the evolutionary basis for genes for resistance to a Eurasian poplar rust fungus, Melampsora larici-populina, that are possessed by the North American species of Populus, P. deltoides (Cervera et al. 1996; Villar et al. 1996), is harder to imagine. We know that M. larici-populina was only introduced to North America in the early 1990s (Newcombe and Chastagner 1993), so the selective force or agent could not have been this fungus. The same question is raised by the abovementioned “Gasaway” gene that confers resistance to a fungus found only in North America even though the gene itself is from a European plant, Corylus avellana. We have already mentioned that species and hybrids of Eucalyptus, introduced to South America, encountered there for the first time a novel rust fungus, Puccinia psidii, which shifted to Eucalyptus from native Myrtaceae (Grgurinovic et al. 2006). Many hybrids of E. grandis of widespread use in Brazilian plantations have proven to be very susceptible to P. psidii; the latter also has a wide host range in the Myrtaceae having been reported on 11 genera and 31 species (Rayachhetry et al. 2001). Nevertheless, there are individuals of E. grandis that are resistant; one harbors a major gene for resistance to P. psidii, the Ppr1 gene (Junghans et al. 2003). How can we explain in terms of selection an Australian gene for resistance to a South American fungus, without going back in time to the Late Paleocene/Early Eocene thermal maximum, 55 mya, when there is evidence for floristic exchange between South America and Australia that included Myrtaceae (Morley 2003)? Moreover, some evolutionarily naive species of Myrtaceae appear resistant, as species, to P. psidii (Rayachhetry et al. 2001). Another example is found in the native, North American range of western gall rust, caused by Endocronartium harknessii. Scots pine (Pinus sylvestris), one of the most widely grown of Eurasian pines in North America, possesses a recessive major gene for resistance 22 G. Newcombe and F.M. Dugan to western gall rust (Van der Kamp 1991). This gene may be common in Scots pine, at least in relation to the population of the western gall rust fungus in British Columbia where the study was performed. Two Asian hard pines, Pinus thunbergii and P. densiflora, are also resistant to western gall rust (Hopkin and Blenis 1989), although genetic analyses of their resistance have not been performed. From Populus deltoides of eastern North America were inherited QTL for resistance to a Pacific Northwestern population of Mycosphaerella populicola (Newcombe and Bradshaw 1996). Asian elms possess genes for resistance to black leaf spot caused by the North American population of Stegophora ulmea (Benet et al. 1995). A last example in this section is that of the NRSA-1 gene for resistance to Striga asiatica that is found in a nonhost, Tagetes erecta, or marigold (Gowda et al. 1999). 1.8 A “Tens Rule” for Novel Pathogens Exapted genes for resistance, such as the Cr, Mmd, Gasaway, and Ppr genes, have been found in resistant individuals in otherwise susceptible species. As such, exapted resistance differs from nonhost resistance in that the latter is presumed to be fixed in species outside the host range of the pathogen in question. But the evolutionary basis for this semantic distinction is unclear. In order to predict outcomes of first encounters between novel pathogens and evolutionarily naive plants, some estimate of frequency is needed, even though the evolutionary basis for the resistant outcome may be unknown. Examples suggest that the frequency of resistant outcomes is probably high. For example, an Asian maple species planted in North America is naive with respect to the “Category 1” pathogens (Table 1.1) of North American native maples. Would Rhytisma americanum infect an Asian maple grown in the U.S.? The answer appears to be no in that R. americanum is limited to North American natives, Acer rubrum and A. saccharinum (Farr et al. n.d.; Hudler and Jensen-Tracy 1998). Asian maples in North America are also apparently resistant to the “Category 2” R. acerinum that occurs on Norway maples, A. platanoides, in North America. The “Category 1” taxa of Mycosphaerellales that are quite common on North American maples in North America also do not appear to attack Asian maples at all (Farr et al. n.d.). Would this pattern hold if we considered a North American plant that has been introduced into a different continent? Consider Pinus contorta, or the North American lodgepole pine, that is utilized quite commonly in forest plantations in northern Europe. Three decades ago, Roll-Hansen noted that P. contorta is “immune or nearly immune to all European rust fungi,” and “more resistant than P. sylvestris to Phacidium infestans and Lophodermium pinastri” (Roll-Hansen 1978). Prunus serotina, or black cherry, provides another good example because it is a North American tree that has become invasive in European forests (Chabrerie et al. 2008). Although eight rust taxa affect Eurasian species of Prunus in Europe, none of these “Category 1” pathogens infect P. serotina (Farr et al. n.d.). In other words, none provide biotic resistance against this plant invader. This is not because 1 Fungal Pathogens of Plants in the Homogocene 23 P. serotina is immune to all rust fungi; in its native range in North America, four rust taxa affect it (Farr et al. n.d.). Similarly, the “Category 2” pear trellis rust, Gymnosporangium fuscum, has remained confined in North America to the Eurasian genus Pyrus as indigenous rosaceous genera are not known to be aecial hosts, and indigenous Juniperus populations are apparently resistant (Ziller 1974). “Category 1” pathogens may switch immediately to long-term, alien plants, or they may eventually produce some virulent propagules that successfully infect the alien. Some past switches were likely not recorded immediately, so the importance of the extended opportunities of a lag period that is used to distinguish categories 2 and 3 (Table 1.1) is not yet clear. For example, Hibiscus syriacus, the popular rose-of-Sharon, was introduced to the Americas in the late sixteenth century from its native range in Asia. In the ensuing, 400 years in the Americas H. syriacus acquired five rust taxa that are not known to occur in its native range (Farr et al. n.d.). But the exact dates of switching are not known. Oddly, the one rust species that does occur on H. syriacus in India, Uromyces heterogeneus, has never been reunited with its host in its introduced range. “Category 2” pathogens are alien pathogens that have been reunited with their adaptive hosts. The latter provide these pathogens with a lag period that they may need to successfully infect naive, native plants. Podosphaera leucotricha causes powdery mildew of apple, Malus domestica, that was domesticated in Eurasia. Like Venturia inaequalis that was shown to be Eurasian in origin (Gladieux et al. 2008) P. leucotricha appears to be Eurasian also. But, P. leucotricha has been reunited with apple in every part of North America in which apples are cultivated (Farr et al. n.d.) such that the seven Malus taxa native to North America have undoubtedly been exposed to its inoculum. Six of the seven appear to be resistant to P. leucotricha in that there are no records of this fungus on them. But one native species of Malus in the southeastern part of the U.S., M. angustifolia, has proven to be susceptible (Table 1.1). Phylogenetic signal does not appear to explain this susceptible exception as M. angustifolia is no more closely related to adaptive host species of Malus than resistant species of North America (e.g., M. coronaria) (Robinson et al. 2001). So, in this case, resistance appears to be exapted rather than nonhost. P. leucotricha has also been reunited in North America with Eurasian species of Photinia, P. glabra and P. serratifolia. Extended opportunities for first encounters with three species of Photinia native to North America have thus also been assured, but outcomes thus far apparently involve nothing but resistance as no records are known. This trend toward resistant outcomes of first encounters continues with two other genera of Rosaceae, Crataegus and Spiraea. Crataegus is especially speciose in North America (USDA n.d.) but there are no records of any of its taxa hosting P. leucotricha, even though C. cuneata in Japan does host P. leucotricha; perhaps this record is of a pathotype that has never been introduced into North America. In the case of Spiraea, P. leucotricha has been recorded on Japanese spiraea, S.  bumalda, in North America, but has not been recorded on 12 taxa of Spiraea native to North America (Farr et al. n.d.). If lag periods do not figure in the outcomes of first encounters, then the three categories of Table 1.1 could be collapsed. 24 G. Newcombe and F.M. Dugan Examples such as these have not been subjected to genetic analysis, but they do suggest a relatively high frequency of resistant outcomes when novel pathogens and naive plants meet. Furthermore, first encounters must be common as naturalized plants in the U.S. belong to 549 genera (USDA n.d.), of which 305, or 56%, are represented in the U.S. by both the naturalized species and native congeners. Other parts of the homogenized world are likely similar in affording many opportunities for plants to encounter “Category 1” or “Category 2” pathogens in particular. A variant of the “tens rule” may thus apply to alien, plant pathogenic fungi in that only a fraction of all first encounters result in susceptible outcomes. This is analogous to the fact that only a fraction of all plant introductions result in plant invasions. This analogy, of course, does not imply that the same mechanism explains both phenomena. Improvements in our knowledge of the “tens rule” for novel fungal pathogens of plants will be built upon advances in the systematics and diagnostics of fungi that allow us to distinguish between pathogen reunions and first encounters. Improvements in our ability to predict which first encounters will result in relatively rare, but devastating, susceptible outcomes will come with a deeper understanding of the evolution and retention of genes for resistance. 1.9 Transformers Susceptible outcomes of novel encounters can however be “transformative” if they change the “character, condition, form or nature of ecosystems over a substantial area” (Pyšek et al. 2004). The chestnut blight fungus, Cryphonectria parasitica, was clearly a “transformer” in the range of Castanea dentata, the American chestnut. The latter was an abundant species in eastern North America at the time of the introduction of the blight fungus (Paillet 2002). C. dentata is no longer a dominant, overstory tree species in those deciduous forests that are starting to be dominated by oak and hickory (McGormick and Platt 1980). Unfortunately, “we know very little concerning ecosystem response to the loss of chestnut” (Orwig 2002). Effects of blight on the food web were probably profound, but they were not studied except anecdotally. The American chestnut itself was not driven to extinction by blight, but chestnut-specific insects likely were (Opler 1978). A particular class of transformer among novel fungal pathogens would be one which does cause the extinction of an evolutionarily naive plant species. However, examples of this are not known. Possibly, novel pathogens came closest with the above-mentioned Franklinia alatamaha. This tree species is not now extinct, but its only natural population was extirpated shortly after the Bartrams discovered it. Were it not for ex situ cultivation, F. alatamaha would now be extinct. Speculation about the causes of the loss of the single, naturally occurring population abounds (Rowland 2006), and that speculation includes the introduction of novel pathogens. Small populations are notoriously susceptible to stochastic forces of extinction that might include pathogens (Rosenzweig 2001b), as we have also briefly 1 Fungal Pathogens of Plants in the Homogocene 25 discussed. This is a serious concern for Wollemia nobilis that lacks genetic variation (Peakall et al. 2003); that finding might indicate that W. nobilis has lost through genetic drift genes for exapted resistance. Little is yet known however of the susceptibilities of W. nobilis other than that it has been shown to be susceptible to Phytophthora cinnamomi and to a species of Botryosphaeria (Bullock et al. 2000). Genetic uniformity certainly affected the magnitude of tree mortality caused by the Dutch elm fungus to Ulmus procera, the English elm, that turned out to be a 2,000year-old Roman clone (Gil et al. 2004). 1.10 Deliberate Introductions of Fungi Deliberate introductions of fungi have likely been uncommon. Some introductions have been made to control plant invaders (i.e., classical biological control) with pathogens with narrow host ranges such as rust fungi (Bruckart and Dowler 1986). Edible, cultivable mushrooms are certainly cultivated outside their native ranges (Arora 1986). Australian ectomycorrhizal fungi “were likely introduced with eucalypt seedlings brought into peninsular Spain before plant quarantine restrictions were observed” (Dı́ez 2005), and this introduction may have been deliberate if the people transporting the seedlings knew of the dependence of eucalypts on these fungi. For the same purpose, ectomycorrhizal associates of pine seedlings were deliberately introduced into the southern hemisphere (Wingfield et al. 2001). Unfortunately, these introductions also inadvertently brought with them soil pathogens of some concern. For example, Rhizina undulata, native to the northern hemisphere, now causes root disease in plantations of northern hemisphere conifers grown in plantations in southern Africa (Wingfield et al. 2001). Armillaria mellea, the root rot fungus, may also have entered South Africa in this way (Coetzee et al. 2001). Another nontarget effect of these ectomycorrhizal introductions, deliberate or inadvertent, has involved competition with native ectomycorrhizal fungi in the exotic tree plantations (Dı́ez 2005). There are no doubt other examples, but in brief summary, deliberately introduced fungi represent just a tiny fraction of global, fungal diversity. 1.11 Inadvertent Co-Introductions of Fungi in Plants Brasier highlights the dangers of inadvertent introductions of fungi by the modern plant trade (Brasier 2008). Even trees “up to 10 m tall with large root balls attached” are being moved from one country to another. Homogenization of previously isolated fungal communities above and belowground is thought to be inevitable if this trade persists. Not only can such shipments not be made safe, but the exotic plant itself contributes to changes in microbial community structure and function in 26 G. Newcombe and F.M. Dugan the soil (Kourtev et al. 2002). As belowground mutualisms (e.g., arbuscular mycorrhizal fungi) (Wolfe et al. 2005) affect aboveground mutualisms (e.g., pollinators) the effects of co-introductions of plants and fungi in the burgeoning plant trade may be profound even if researchers have not yet elucidated all of them. The movement of fungi in international food shipments can only be guessed at. Pathogen release might seem almost miraculous when considered in the light of the ease with which pathogens have sometimes been moved with their host. For example, decades ago, Savile noted the inconspicuous adherence of teliospores of Puccinia carthami to seeds of safflower that when “planted in an isolated garden produced seedlings with pycnia” (Savile 1973). How then in 1876 did Henry Wickham succeed in moving 70,000 seeds of rubber-producing Hevea brasiliensis from its native Amazon basin to the Old World tropics via Kew without any propagules of the notorious blight fungus, Microcyclus ulei (Hobhouse 2003)? Wickham’s move and the subsequent pathogen release enjoyed by rubber plantations in Southeast Asia changed the course of the twentieth century. And if Henry Ford had understood pathogen release better, he might have thought twice about trying to duplicate the success of Asian rubber plantations by attempting to establish in the 1920s plantations in the Amazon that failed miserably due to blight. Even if seeds and other plant propagules are surface-sterilized, endophytes are still moved around the world in the plant trade (Palm 1999). Endophytes can affect the ecology of plants in many ways, from tolerance to stressful conditions (Redman et al. 2002), through growth effects (Ernst et al. 2003), to plant community diversity (Clay and Holah 1999). We are still filling in the knowledge gaps of the functional roles in what has been called the “endophytic continuum” (Schulz and Boyle 2005). Endophytes in one invasive plant, Centaurea stoebe or spotted knapweed, were recently reported to be remarkably diverse in both the native and invaded ranges of the plant (Shipunov et al. 2008), with interesting effects on its ecology (Newcombe et al. 2009). Analyses of these communities suggested that both host switching and co-introduction “took place during the knapweed invasion.” It is possible that the origins of such fungi as the Dutch elm disease and dogwood anthracnose pathogens have not been determined because these fungi are only endophytic in the native ranges of their hosts. 1.12 Fungi as Facilitators of Plant Invasions Release from phytopathogenic fungi is but one hypothesis to explain plant invasions (Mitchell and Power 2003). Recently, plant invasion biologists have taken a new interest in mycorrhization and other mutualisms (Richardson et al. 2000a), and in plant–soil feedback processes more generally (Ehrenfeld et al. 2005). Fungi in soil appear to be central to plant–soil feedbacks that promote alien plant invasions (Klironomos 2002). Alien plants may become abundant in part because they are 1 Fungal Pathogens of Plants in the Homogocene 27 relatively resistant to fungal pathogens in soil that limit the abundance of many native plants (Klironomos 2002). Interpreted in light of the foregoing discussion of a “tens rule” for novel plant pathogens this would again indicate that first encounters below ground of novel pathogens and naive plants may more often than not involve incompatibility or resistance, just as aboveground encounters do. This point is reinforced by a meta-analysis of biotic resistance that revealed that native communities are defended against plant invaders by resident competitors and herbivores, but not by soil fungi (Levine et al. 2004). In other words, an improved understanding of the outcomes of first encounters between novel pathogens and naive plants can potentially contribute simultaneously to questions that have been treated separately by different disciplines. On the one hand, mycologists and plant pathologists have had a traditional interest in predicting the outcome of introductions of novel plant pathogens. On the other, plant ecologists and invasion biologists have been trying to understand the mechanism of plant invasions. When the fungal pathogens in a plant community are native (i.e., Category 1 of Table 1.1), they fail to contribute to biotic resistance to plant invasions insofar as they fail to cause disease of alien plants, as summarized by Levine. This result is consistent with examples in categories 2 and 3 where resistant outcomes again prevail. 1.13 Conclusions Ideally, on the eve of the Homogocene in 1499, trained scientists around the globe would have already described all species of life. A pre-Homogocene “Encyclopedia of Life” (Wilson 2003) would have provided the baseline from “Day 1” for subsequent tracking of every human-aided introduction of an alien that followed. Knowing the outcomes of every introduction of an alien species during the past 500 years, we might well be further along in our attempt to understand why some organisms are invasive and why most are not. Unfortunately, 1500 predates the development of Linnean taxonomy by 253 years (Linnaeus 1753), Christiaan H. Persoon’s Synopsis Methodica Fungorum by 301 years, and Elias Magnus Fries’ first volume of Systema Fungorum by 321 years. Falling far short of the ideal, we instead find ourselves drawing inferences from patterns of pathogen release, and scattered studies of resistance relative to first encounters of plants and pathogens. Fortunately, for inferences from pathogen release, the SMML Fungus–Host Distribution Database and other databases of records of fungi on plants have proven invaluable. Homogenization itself has created opportunities for discoveries that otherwise might not have been made. Finally, the merging of the study of novel pathogens of plants with the general framework and terminology of invasion biology is bound to be helpful to all students of this global experiment of the last 500 years. 28 G. Newcombe and F.M. Dugan References Abbo S, Shtienberg D, Lichtenzveig J, Lev-Yadun S, Gopher A (2003) The chickpea, summer cropping, and a new model for pulse domestication in the ancient near east. Quart Rev Biol 78:435–448 Amin S (1974) Accumulation on a world scale. Monthly Review Press, New York Armbruster WS (1997) Exaptations link evolution of plant-herbivore and plant-pollinator interactions: a phylogenetic inquiry. Ecology 78:1661–1672 Arora D (1986) Mushrooms demystified, 2nd edn. Ten Speed Press, Berkeley, CA Baskin Y (2003) A plague of rats and rubbervines. Island Press, Washington, D.C Behre KE (1992) The history of rye cultivation in Europe. Veg Hist Archaeobot 1:141–156 Benet H, Guries RP, Boury S, Smalley EB (1995) Identification of RAPD markers linked to a black leaf spot resistance gene in Chinese elm. Theor Appl Genet 90:1068–1073 Bishop EM (2003) In search of ancient Oregon: a geological and natural history. Timber Press, Inc., Portland Brasier C (2000) The rise of the hybrid fungi. Nature 405:134–135 Brasier CM (1990) China and the origins of Dutch elm disease: an appraisal. Plant Pathol 39:5–16 Brasier CM (2008) The biosecurity threat to the UK and global environment from international trade in plants. Plant Pathol 57:792–808 Brasier CM, Buck KW (2001) Rapid evolutionary changes in a globally invading fungal pathogen (Dutch elm disease). Biol Invasions 3:223–233 Brasier CM, Mehrotra M (1995) Ophiostoma himal-ulmi sp. nov., a new species of Dutch elm disease fungus endemic to the Himalayas. Mycol Res 99:205–215 Briggs SP, Johal GS (1994) Genetic patterns of plant host-parasite interactions. Trends Genet 10:12–16 Bruckart WL, Dowler WM (1986) Evaluation of exotic rust fungi in the United States for classical biological control of weeds. Weed Sci 34:11–14 Bullock S, Summerell BA, Gunn LV (2000) Pathogenicity of five fungi to the Wollemi pine, Wollemia nobilis. Australas Plant Pathol 29:211–214 Cain RF (1972) Evolution of the fungi. Mycologia 64:1–14 Callaway R, Ridenour WM (2004) Novel weapons: invasive success and the evolution of increased competitive ability. Front Ecol Environ 2:436–443 Campbell BMS (1988) The diffusion of vetches in medieval England. Econ Hist Rev 41:193–208 Carlquist S (1980) Hawaii: a natural history, 2nd edn. Pacific Tropical Botanical Garden, Lawai, Kaua’i Cervera MT et al (1996) Identification of AFLP molecular markers for resistance against Melampsora larici-populina in Populus. Theor Appl Genet 93:733–737 Chabrerie O, Verheyen K, Saguez R, Decocq G (2008) Disentangling relationships between habitat conditions, disturbance history, plant diversity, and American black cherry (Prunus serotina Ehrh.) invasion in a European temperate forest. Divers Distrib 14:204–212 Clay K, Holah J (1999) Fungal endophyte symbiosis and plant diversity in successional fields. Science 285:1742–1744 Coetzee M, Wingfield BD, Harrington TC, Steimel J, Coutinho TA, Wingfield MJ (2001) The root rot fungus Armillaria mellea introduced into South Africa by early Dutch settlers. Mol Ecol 10:387–396 Cox CB, Moore PD (2005) Biogeography – an ecological and evolutionary approach, 7th edn. Blackwell, Oxford, U.K Cox GW (1999) Alien species in North America and Hawaii. Island Press, Washington, D.C Coyne CJ, Mehlenbacher SA, Smith DC (1998) Sources of resistance to eastern filbert blight in hazelnut. J Am Soc Hort Sci 123:253–257 Critchfield WB (1985) The late Quaternary history of lodgepole and jack pines. Can J Forest Res 15:749–772 1 Fungal Pathogens of Plants in the Homogocene 29 Crosby AW (2004) Ecological Imperialism: the biological expansion of Europe, 900–1900. Cambridge University Press, Cambridge, UK Cullen JM, Kable PF, Catt M (1973) Epidemic spread of a rust imported for biological control. Nature 244:462–464 Dark P, Gent H (2001) Pests and diseases of prehistoric crops: a yield “honeymoon” for early grain crops in Europe. Oxford J Archaeol 20:59–78 Darwin C (1859) On the origin of species. Murray, London Diamond J, Bellwood P (2003) Farmers and their languages: the first expansions. Science 300:597–603 Dı́ez J (2005) Invasion biology of Australian ectomycorrhizal fungi introduced with eucalypt plantations into the Iberian Peninsula. Biol Invasions 7:3–15 Dugan F (2008) Fungi in the ancient World: how mushrooms, mildews, molds and yeast shaped the early civilizations of Europe, the Mediterranean, and the near east. APS Press, St. Paul, MN Dugan FM, Carris LM (1992) Puccinia jaceae var. diffusa and P. acroptili on knapweeds in Washington. Plant Dis 76:972 Duncan WH, Duncan MB (1988) Trees of the Southeastern United States. University of Georgia Press, Athens, Georgia Eckenwalder JE (1984) Natural intersectional hybridization between North American species of Populus (Salicaceae) in sections Aigeiros and Tacamahaca. III. Paleobotany and evolution. Can J Bot 62:336–342 Eckenwalder JE (1996) Systematics and evolution of Populus. Part I, Chap. 1. In: Stettler RF, Bradshaw HD Jr, Heilman PE, Hinckley TM (eds) Biology of Populus and its implications for management and conservation. NRC Research Press, National Research Council of Canada, Ottawa, ON, pp 7–32 Ehrenfeld JG, Ravit B, Elgersma K (2005) Feedback in the plant-soil system. Annu Rev Environ Resour 30:75–115 Ellstrand NC, Schierenbeck K (2000) Hybridization as a stimulus for the evolution of invasiveness in plants. PNAS 97:7043–7050 Ernst M, Mendgen KW, Wirsel SGR (2003) Endophytic fungal mutualists: seed-borne Stagonospora spp. enhance reed biomass production in axenic microcosms. Mol Plant-Microbe Interact 16:580–587 Farr DF, Rossman AY, Palm ME, McCray EB (n.d.) Fungal Databases. Systematic Mycology and Microbiology, USDA, ARS http://nt.ars-grin.gov/fungaldatabases/ Fenchel T, Finlay BJ (2004) The ubiquity of small species: patterns of local and global diversity. Bioscience 54:777–784 Flor HH (1971) Current status of the gene-for-gene concept. Annu Rev Phytopathol 9:275–296 Genton BJ, Shykoff JA, Giraud T (2005) High genetic diversity in French invasive populations of common ragweed, Ambrosia artemisiifolia, as a result of multiple sources of introduction. Mol Ecol 14:4275–4285 Gernandt DS, Geada López GG, Ortiz Garcı́a S, Liston A (2005) Phylogeny and classification of Pinus. Taxon 54:29–42 Gil L, Fuentes-Utrilla P, Soto A, Cervera MT, Collada C (2004) English elm is a 2, 000-year-old Roman clone. Nature 431:1053 Gilbert GS, Webb CO (2007) Phylogenetic signal in plant pathogen-host range. Proc Natl Acad Sci 104:4979–4983 Gladieux P, Zhang X-G, Afoufa-Bastien D, Valdebenito Sanhueza R-M, Sbaghi M, Le Cam B (2008) On the origin and spread of the scab disease of apple: out of central Asia. PLoS ONE 3:1–14 Gómez-Alpizar L, Carbone I, Ristaino JB (2007) An Andean origin of Phytophthora infestans inferred from mitochondrial and nuclear gene genealogies. PNAS 104:3306–3311 Gould SJ, Vrba ES (1982) Exaptation – a missing term in the science of form. Paleobiology 8:4–15 Gowda BS, Riopel JL, Timko MP (1999) NRSA-1: a resistance gene homolog expressed in roots of non-host plants following parasitism by Striga asiatica (witchweed). Plant J 20:217–230 30 G. Newcombe and F.M. Dugan Gräser Y, Fröhlich J, Presber W, De Hoog S (2007) Microsatellite markers reveal geographic population differentiation in Trichophyton rubrum. J Med Microbiol 56:1058–1065 Green JL et al (2004) Spatial scaling of microbial eukaryote diversity. Nature 432:747–750 Grgurinovic CA, Walsh D, Macbeth F (2006) Eucalyptus rust caused by Puccinia psidii and the threat it poses to Australia. EPPO Bull 36:486–489 Grove WB (1913) The British rust fungi (Uredinales): their biology and classification. Cambridge University Press, Cambridge Hafner MS, Nadler SA (1988) Phylogenetic trees support the coevolution of parasites and their hosts. Nature 332:258–259 Harvey JH (1984) Vegetables in the middle ages. Garden Hist 12:89–99 Harvey JH (1992) Garden plants of Moorish Spain: a fresh look. Garden History 20:71–82 Haughton CS (1978) Green immigrants: the Plants that Transformed America. Harcourt Brace Jovanovich, New York Hawksworth DL (2001) The magnitude of fungal diversity: the 1.5 million species estimate revisited. Mycol Res 105:1422–1432 Hedrick UP (1950) A history of horticulture in America to 1860. Oxford University Press, New York, NY Heger T, Trepl L (2003) Predicting biological invasions. Biol Invasions 5:313–321 Hobhouse H (2003) Seeds of wealth. Shoemaker & Hoard, Washington, D.C Holdenrieder O, Sieber TN (2007) First record of Discula destructiva in Switzerland and preliminary inoculation experiments on native European Cornus species. In: Evans H, Oszako T (eds) Alien invasive species and International trade. Forest Research Institute, Warsaw, Poland, pp 51–56 Hopkin AA, Blenis PV (1989) Resistant responses in juvenile seedlings of Pinus densiflora (Japanese red pine) inoculated with Endocronartium harknessii. Can J Bot 67:3545–3552 Hudler GW, Jensen-Tracy S (1998) Rhytisma americanum sp. nov.: a previously undescribed species of Rhytisma on maples (Acer spp.). Mycotaxon 68:405–416 Jackson AP (2004) A reconciliation analysis of host switching in plant-fungal symbioses. Evolution 58:1909–1923 Jones DR, Baker RHA (2007) Introductions of non-native plant pathogens into Great Britain, 1970–2004. Plant Pathol 56:891–910 Junghans DT, Alfenas AC, Brommonschenkel SH, Oda S, Mello EJ, Grattapaglia D (2003) Resistance to rust (Puccinia psidii Winter) in Eucalyptus: mode of inheritance and mapping of a major gene with RAPD markers. Theor Appl Genet 108:175–180 Juzwik J, Harrington TC, MacDonald WL, Appel DN (2008) The origin of Ceratocystis fagacearum, the oak wilt fungus. Annu Rev Phytopathol 46:13–26 Kasanen R, Hantula J, Ostry ME, Pinon J, Kurkela T (2004) North American populations of Entoleuca mammata are genetically more variable than populations in Europe. Mycol Res 108:766–774 Kauserud H et al (2007) Asian origin and rapid global spread of the destructive dry rot fungus Serpula lacrymans. Mol Ecol 16:3350–3360 Keane RM, Crawley MJ (2002) Exotic plant invasions and the enemy release hypothesis. Trends Ecol Evol 17:164–170 Kinloch BB (1992) Distribution and frequency of a gene for resistance to white pine blister rust in natural populations of sugar pine. Can J Bot 70:1319–1323 Kinloch BB (2003) White pine blister rust in North America: past and prognosis. Phytopathology 93:1044–1047 Kinloch BB, Dupper GE (2002) Genetic specificity in the white pine-blister rust pathosystem. Phytopathology 92:278–280 Kinloch BB, Sniezko RA, Barnes GD, Greathouse TE (1999) A major gene for resistance to white pine blister rust in western white pine from the western Cascade Range. Phytopathology 89:861–867 Kinloch BB, Sniezko RA, Dupper GE (2003) Origin and distribution of Cr2, a gene for resistance to white pine blister rust in natural populations of western white pine. Phytopathology 93:691–694 1 Fungal Pathogens of Plants in the Homogocene 31 Klironomos JN (2002) Feedback with soil biota contributes to plant rarity and invasiveness in communities. Nature 417:67–70 Kolar CS, Lodge DM (2001) Progress in invasion biology: predicting invaders. Trends Ecol Evol 16:199–204 Kourtev PS, Ehrenfeld JG, Haggblom M (2002) Exotic plant species alter the microbial community structure and function in the soil. Ecology 83:3152–3166 Kroll H (2005) Literature on archaeological remains of cultivated plants 1981–2004. In. www. archaeobotany.de/database.html Levine JM, Adler PB, Yelenik SG (2004) A meta-analysis of biotic resistance to exotic plant invasions. Ecol Lett 7:975–989 Liebhold AM, MacDonald WL, Bergdahl D, Mastro VC (1995) Invasion by exotic forest pests: a threat to forest ecosystems. Forest Sci Monograph 30:1–57 Linnaeus C (1753) Species plantarum. Laurentius Salvius, Sweden Love RS (2006) Maritime exploration in the age of discovery, 1415–1800 Greenwood Press Mabberley DJ (2008) Mabberley’s plant book, 3rd edn. Cambridge University Press, New York Mack RN, Simberloff D, Lonsdale WM, Evans H, Clout M, Bazzaz FA (2000) Biotic invasions: causes, epidemiology, global consequences, and control. Ecol Appl 10:689–710 Madden LV (2001) What are the nonindigenous plant pathogens that threaten US crops and forests? APSnet Feature Story: http://www. apsnet. org/online/feature/exotic Maddox DM (1982) Biological control of diffuse knapweed (Centaurea diffusa) and spotted knapweed (C. maculosa). Weed Sci 30:76–82 Marı́n DH, Romero RA, Guzmán M, Sutton TB (2003) Black sigatoka: an increasing threat to banana cultivation. Plant Dis 87:208–222 McGormick JF, Platt RB (1980) Recovery of an Appalachian forest following the chestnut blight or Catherine Keever – you were right. Am Midl Nat 104:264–273 Mitchell CE, Power AG (2003) Release of invasive plants from fungal and viral pathogens. Nature 421:625–627 Mooney HA, Cleland EE (2001) The evolutionary impact of invasive species. Proceedings of the National Academy of Sciences 98:5446–5451 Morley RJ (2003) Interplate dispersal paths for megathermal angiosperms. Perspect Plant Ecol Evol Syst 6:5–20 Mortensen K, Harris P, Makowski RMD (1989) First occurrence of Puccinia jaceae var. diffusae in North America on diffuse knapweed (Centaurea diffusa). Can J Plant Pathol 11:322–324 Newcombe G (1998) A review of exapted resistance to diseases of Populus. Eur J Forest Pathol 28:209–216 Newcombe G (2003a) Native Venturia inopina sp. nov., specific to Populus trichocarpa and its hybrids. Mycol Res 107:108–116 Newcombe G (2003b) Puccinia tanaceti – specialist or generalist. Mycol Res 107:797–802 Newcombe G (2005) Genes for parasite-specific, nonhost resistance in Populus. Phytopathology 95:779–783 Newcombe G, Bradshaw HD Jr (1996) Quantitative trait loci conferring resistance in hybrid poplar to Septoria populicola, the cause of leaf spot. Can J For Res 26:1943–1950 Newcombe G, Bradshaw HD Jr, Chastagner GA, Stettler RF (1996) A major gene for resistance to Melampsora medusae f. sp. deltoidae in a hybrid poplar pedigree. Phytopathology 86:87–94 Newcombe G, Chastagner GA (1993) First report of the Eurasian poplar leaf rust fungus, Melampsora laricici-populina, in North America. Plant Dis 77:532–535 Newcombe G et al (2009) Endophytes influence protection and growth of an invasive plant. Commun Integr Biol 2:1–3 Newcombe G, Stirling B, Bradshaw HD Jr (2001) Abundant pathogenic variation in the new hybrid rust Melampsora x columbiana on hybrid poplar. Phytopathology 91:981–985 Newcombe G, Stirling B, McDonald S, Bradshaw HD (2000) Melampsora x columbiana, a natural hybrid of M. medusae and M. occidentalis. Mycol Res 104:261–274 32 G. Newcombe and F.M. Dugan Novak SJ, Mack RN (2005) Genetic bottlenecks in alien plant species. In: Sax DF, Stachowicz JJ, Gaines SD (eds) Species invasions: insights into ecology, evolution, and biogeography. Sinauer Sunderland, Massachusetts, pp 201–228 Olden JD, Poff NL, Douglas MR, Douglas ME, Fausch KD (2004) Ecological and evolutionary consequences of biotic homogenization. Trends Ecol Evol 19:18–24 Opler PA (1978) Insects of American chestnut: possible importance and conservation concerns. In: McDonald J (ed) The American chestnut symposium. West Virginia University Press, Morgantown, West Virginia Orwig DA (2002) Ecosystem to regional impacts of introduced pests and pathogens: historical context, questions and issues. J Biogeogr 29:1471–1474 Ostry ME (1997) Sirococcus clavigignenti-juglandacaerum on heartnut (Juglans ailantifolia var. cordiformis). Plant Dis 81:1461 Ostry ME, Moore M (2007) Natural and experimental host range of Sirococcus clavigignentijuglandacearum. Plant Dis 91:581–584 Ouborg NJ, Vergeer P, Mix C (2006) The rough edges of the conservation genetics paradigm for plants. J Ecol 94:1233–1248 Paillet FL (2002) Chestnut: history and ecology of a transformed species. J Biogeogr 29: 1517–1530 Palm ME (1999) Mycology and world trade: a view from the front line. Mycologia 91:1–12 Palm ME (2001) Systematics and the impact of invasive fungi on agriculture in the United States. Bioscience 51:141–147 Palm ME, Richard RD, Parker P (1992) First report of Puccinia jaceae var. diffusa on diffuse knapweed in the United States. Plant Dis 76:972 Palm ME, Rossman AY (2003) Invasion pathways of terrestrial plant-inhabiting fungi. In: Ruiz GM, Carlton JT (eds) Invasive species – vectors and management strategies. Island Press, Washington, D.C, p 518 Parker IM, Gilbert GS (2004) The evolutionary ecology of novel plant-pathogen interactions. Annu Rev Ecol Evol Syst 35:675–700 Parmesan C (2006) Ecological and evolutionary responses to recent climate change. Annu Rev Ecol Evol Syst 37:637–669 Peakall R, Ebert D, Scott L, Meagher P, Offord C (2003) Comparative genetic study confirms exceptionally low genetic variation in the ancient and endangered relictual conifer, Wollemia nobilis (Araucariaceae). Mol Ecol 12:2331–2343 Person C (1959) Gene-for-gene relationships in host: parasite systems. Can J Bot 37:1101–1130 Person C (1967) Genetic aspects of parasitism. Can J Bot 45:1193–1204 Pimentel D, Zuniga R, Morrison D (2005) Update on the environmental and economic costs associated with alien-invasive species in the United States. Ecol Econ 52:273–288 Preston CD, Pearman D, Hall A (2004) Archaeophytes in Britain. Bot J Linn Soc 145:257–294 Pyšek P, Richardson DM, Jarošı́k V (2006) Who cites who in the invasion zoo: insights from an analysis of the most highly cited papers in invasion ecology. Preslia 78:437–468 Pyšek P, Richardson DM, Rejmánek M, Webster GL, Williamson M, Kirschner J (2004) Alien plants in checklists and floras: towards better communication between taxonomists and ecologists. Taxon 53:131–143 Rayachhetry MB, Van TK, Center TD, Elliott ML (2001) Host range of Puccinia psidii, a potential biological control agent of Melaleuca quinquenervia in Florida. Biol Control 22:38–45 Rebourg C, Chastanet M, Gousenard B, Welcker C, Dubreuil P, Charcosset A (2003) Maize introduction into Europe: the history reviewed in the light of molecular data. Theor Appl Genet 106:895–903 Redlin SC (1991) Discula destructiva sp. nov., cause of dogwood anthracnose. Mycologia 83:633–642 Redman RS, Seehan KB, Stout RG, Rodriquez RJ, Henson JM (2002) Thermotolerance generated by plant/fungal symbiosis. Science 298:1581 Rehder A (1940) Manual of cultivated trees and shrubs, 2nd edn. MacMillan, New York 1 Fungal Pathogens of Plants in the Homogocene 33 Reichard SH, White P (2001) Horticulture as a pathway of invasive plant introductions in the United States. Bioscience 51:103–113 Rejmánek M (1996) A theory of seed plant invasiveness: the first sketch. Biol Conserv 78:171–181 Richardson DM, Allsopp N, D’Antonio CM, Milton SJ, Rejmanek M (2000a) Plant invasions – the role of mutualisms. Biol Rev 75:65–93 Richardson DM, Pyšek P, Rejmánek M, Barbour MG, Panetta FD, West CJ (2000b) Naturalization and invasion of alien plants: concepts and definitions. Divers Distrib 6:93–107 Rizzo DM, Garbelotto M (2003) Sudden oak death: endangering California and Oregon forest ecosystems. Front Ecol Environ 1:197–204 Robinson JP, Harris SA, Juniper BE (2001) Taxonomy of the genus Malus Mill. (Rosaceae) with emphasis on the cultivated apple, Malus domestica Borkh. Plant Syst Evol 226:35–58 Roll-Hansen F (1978) Fungi dangerous at Pinus contorta with special reference to pathogens from North Europe. Eur J Forest Pathol 8:1–14 Rosenzweig ML (1995) Species diversity in space and time. Cambridge University Press, Cambridge, UK Rosenzweig ML (2001a) The four questions: what does the introduction of exotic species do to diversity. Evol Ecol Res 3:361–367 Rosenzweig ML (2001b) Loss of speciation rate will impoverish future diversity. Proc Natl Acad Sci 98:5404–5410 Rossman AY (2001) A special issue on global movement of invasive plants and fungi. Bioscience 51:93–94 Rossman AY (2009) The impact of invasive fungi on agricultural ecosystems in the United States. Biol Invasions 11:97–107 Rowland LM (2006) America’s first rare plant: the Franklin tree. Terrain Org 18:183–190 Roy BA, Kirchner JW (2000) Evolutionary dynamics of pathogen resistance and tolerance. Evolution 54:51–63 Savile DBO (1970a) Autoecious Puccinia species attacking Cardueae in North America. Can J Bot 48:1567–1584 Savile DBO (1970b) Some Eurasian Puccinia species attacking Cardueae. Can J Bot 48:1553–1566 Savile DBO (1973) Rusts that pass import inspection. Can Plant Dis Surv 53:105–106 Schubert K et al (2007) Biodiversity in the Cladosporium herbarum complex (Davidiellaceae, Capnodiales), with standardisation of methods for Cladosporium taxonomy and diagnostics. Stud Mycol 58:105–156 Schulz B, Boyle C (2005) The endophytic continuum. Mycol Res 109:661–686 Shipunov A, Newcombe G, Raghavendra A, Anderson C (2008) Hidden diversity of endophytic fungi in an invasive plant. Am J Bot 95:1096–1108 Sinclair WA, Lyon HH (2005) Diseases of trees and shrubs, 2nd edn. Cornell University Press, Ithaca, NY Sinclair WA, Lyon HH, Johnson WT (1987) Diseases of trees and shrubs. Cornell University Press, Ithaca, NY Smith H, Wingfield MJ, De Wet J, Coutinho TA (2000) Genotypic diversity of Sphaeropsis sapinea from South Africa and Northern Sumatra. Plant Dis 84:139–142 Smith PM (1986) Native or introduced? problems in the taxonomy and plant geography of some widely introduced annual brome-grasses. Proc R Soc Edinb 89B:273–281 Sniezko RA, Kegley AJ, Danchok R (2008) White pine blister rust resistance in North American, Asian and European species – results from artificial inoculation trials in Oregon. Ann For Res 51:53–66 Stephan BR (2001) Studies of genetic variation with five-needle pines in Germany. In: Sniezko RA (ed) Breeding and genetic resources of five-needle pines: growth, adaptability and pest resistance, vol. RMRS-P-32. USDA Forest Service, Medford, USA, pp 98–102 Stubblefield SP, Taylor TN (1988) Recent advances in palaeomycology. New Phytol 108:3–25 Stukenbrock EH, McDonald BA (2008) The origins of plant pathogens in agro-ecosystems. Annu Rev Phytopathol 46:75–100 34 G. Newcombe and F.M. Dugan Ta CH, Bartholomew B (1984) Camellias. Timber Press, Portland, Oregon Taavitsainen J-P, Simola H, Grönlund E (1998) Cultivation history beyond the periphery: early agriculture in the north European boreal forest. J World Prehist 12:199–253 Tainter FH, Anderson RL (1993) Twenty-six new pine hosts of fusiform rust. Plant Dis 77:17–20 Taylor JW, Turner E, Townsend JP, Dettman JR, Jacobson D (2006) Eukaryotic microbes, species recognition and the geographic limits of species: examples from the kingdom fungi. Philos Trans R Soc Lond B Biol Sci 361:1947–1963 Thompson JN, Burdon JJ (1992) Gene-for-gene coevolution between plants and parasites. Nature 360:121–125 Tuskan GA et al (2006) The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science 313:1596–1604 USDA, ARS, National Genetic Resources Program (n.d.) Germplasm Resources Information Network – (GRIN) [Online Database]. National Germplasm Resources Laboratory, Beltsville, MD Van der Kamp BJ (1991) Major gene resistance of Scots pine to western gall rust. Can J For Res 21:375–378 Van Der Plank JE (1975) Principles of plant infection. Academic Press, New York, NY Vaughan DA, Balázs E, Heslop-Harrison JS (2007) From crop domestication to super-domestication. Ann Bot 100:893–901 Villar M, Lefevre F, Bradshaw HD, Teissier du Cros E (1996) Molecular genetics of rust resistance in poplars (Melampsora larici-populina Kleb/Populus sp.) by bulked segregant analysis in a 2  2 factorial mating design. Genetics 143:531–536 Wallerstein E (1974) The modern World-system, vol I: capitalist agriculture and the origins of the European World-economy in the sixteenth century. Academic Press, New York/London Wallerstein E (1980) The modern World-system, vol II: mercantilism and the consolidation of the European World-economy 1600–1750. Academic Press, New York Wallerstein E (1989) The modern World-system, vol. III: the second great expansion of the capitalist World-economy, 1730–1840’s. Academic Press, San Diego Ward SM, Gaskin JF, Wilson LM (2008) Ecological genetics of plant invasion: what do we know. Invasive Plant Sci Manag 1:98–109 Weldon C, Du Preez LH, Hyatt AD, Muller R, Speare R (2004) Origin of the amphibian chytrid fungus. Emerging Infect Dis 10:2100–2105 Wells S (2007) Deep ancestry: inside the genographic project: the landmark DNA quest to decipher our distant past. National Geographic Society, Washington, D.C Williamson M, Fitter A (1996) The varying success of invaders. Ecology 77:1661–1666 Wilson EO (2003) The encyclopedia of life. Trends Ecol Evol 18:77–80 Wingfield MJ, Slippers B, Roux J, Wingfield BD (2001) Worldwide movement of exotic forest fungi, especially in the tropics and the southern hemisphere. Bioscience 51:134–140 Wolfe BE, Husband BC, Klironomos JN (2005) Effects of a belowground mutualism on an aboveground mutualism. Ecol Lett 8:218–223 Wu HX, Ying CC, Muir JA (1996) Effect of geographic variation and jack pine introgression on disease and insect resistance in lodgepole pine. Can J For Res 26:711–726 Wyand RA, Brown JKM (2003) Genetic and forma specialis diversity in Blumeria graminis of cereals and its implications for host-pathogen co-evolution. Mol Plant Pathol 4:187–198 Zaffarano PL, McDonald BA, Zala M, Linde CC (2006) Global hierarchical gene diversity analylis suggests the fertile crescent is not the center of origin of the barley scald pathogen Rhyncosporium secalis. Phytopathology 96:941–950 Zambino PJ, Harrington TC (2005) Ceratocystis fagacearum: where did it come from? clues from genetic diversity and relatives. Phytopathology 95:S126 Zerbe S, Wirth P (2006) Non-indigenous plant species and their ecological range in Central European pine (Pinus sylvestris L.) forests. Ann For Sci 63:189–203 Ziller WG (1974) The tree rusts of Western Canada. Canadian Forestry Service, Victoria, British Columbia Chapter 2 Molecular Techniques for Classification and Diagnosis of Plant Pathogenic Oomycota Otmar Spring and Marco Thines Abstract With a delay of approximately 10 years, molecular techniques came in use for the investigation of phylogenetic, taxonomic, and diagnostic problems in oomycetes. The particular problem in plant pathogenic Oomycota lies in their biotrophic nature, which prohibits axenic cultivation of the majority of species, in particular downy mildews and white blister rusts, on artificial media. This impeded the broad employment of basic techniques such as RFLP (restriction fragment length polymorphism) in investigations of Oomycota and required the development of specific PCR-based tools for identification and detection of minute pathogen amounts. When the first sequence analysis of genomic loci of oomycetes was conducted in the late 1980s, specific primers became available which allowed selective analysis of oomycete DNA in the presence of much higher amounts of host DNA. Since about 8 years, these methods have become routine in this field of research and have started turning the systematics of Oomycota upside down. A wide array of tools for the amplification of coding and noncoding gene loci helped to differentiate pathogen accessions, to restructure the phylogeny, to form monophyletic entities on all taxonomic levels, and to resolve unrealistically broad species concepts. Recent progress in sequencing ancient DNA from Oomycota allows the extension of taxonomic studies to herbarium collections. This broadens the basis of samples considerably and gives the chance to link molecular phylogenetic taxonomy with the traditional phenotype-based system. Moreover, molecular techniques gain growing importance in the identification of downy mildews and white blister rusts in plant pathology and in ecological studies. Their employment allows detection of Oomycota in asymptomatic infections of host plants and enables the identification of infested seeds or soils in agriculture. With the first whole genome sequencing of a Phytophthora species in 2003, the basis for functional genomics studies has been established. This will not only stimulate phylogenetic, taxonomic, O. Spring and M. Thines Institute of Botany, University of Hohenheim, 70593 Stuttgart, Germany e-mail: spring@uni-hohenheim.de Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_2, # Springer-Verlag Berlin Heidelberg 2010 35 36 O. Spring and M. Thines and diagnostic research in plant pathogenic Oomycota, but also provide the basis for gaining deeper insights in their biology and interaction with their hosts. 2.1 Introduction Molecular techniques have altered the understanding of the evolution and phylogeny of life within the past 30 years in an unprecedented dimension. It is only natural that this process started with the most familiar and the easiest accessible groups of organisms such as animals and higher plants. These investigations revealed a multitude of inconsistencies of the evolutionary relationships compared to the established taxonomy, especially above the generic level. The consequences for the reorganization of the systematics and taxonomy of these already well-known groups were however moderate and manageable, when compared with less-familiar groups such as prokaryotes, protozoa, algae, or fungi. The taxonomic rearrangement of the fungi started long before molecular techniques were at hand. The separation in Myxomycetes, Eumycetes, and Oomycota was based on fundamental cytological and biochemical differences. The particular mode of sexual reproduction of the diploid Oomycota is unparalleled in any other fungus-like entity (Tommerup 1981). In addition, their cellulosic cell wall (Bartnicki-Garcia 1968), the lysine synthesis pathway (Vogel 1960), and their sterol biosynthesis (Warner et al. 1982) are biochemical evidence to separate them from eumycotic fungi, to whom they resemble only superficially with respect to their hyphoidal organization and osmotrophic nutrition. The phylogenetic roots of the Oomycota became obvious from the flagellum, which because of the anterior insertion and its tripartite hairs unraveled them as part of the Straminipila (Vlk 1939; Patterson 1989; Dick 2001). Molecular data supported this origin (Leipe et al. 1994) and hinted at the possibility that Oomycota had separated before or after the uptake of secondary chloroplasts in this clade. Sequencing of the nucleus-encoded glyceraldehyde-3phosphate dehydrogenase in Oomycota supports close relationships with the photosynthetic heterokonts in which this enzyme is plastid-targeted, thus favoring the loss of plastids in the nonphotosynthetic heterokonts (Harper and Keeling 2003). However, the initial rearrangements of the higher-level classification of the fungal organisms gave but a glimpse of the changes to come on all taxonomic levels with the advent of molecular phylogenetic investigations. Within the Oomycota, until recently, the taxonomic concept was solely based on the few morphological characters available from the sparse features of an unsepted mycelium, sexually or asexually produced spores, and the mode of sporulation. In biotrophic sections of this phylum, additional information could be gained from host specificity, provided that vital material was accessible and the assumptions from field observation could be ensured by infection experiments. The host-based concept for species was broadly applied by Gäumann (1918, 1923) and still predominates in taxonomy. Paucity in differentiating morphological characters, in 2 Molecular Techniques for Classification and Diagnosis 37 addition to the lack of knowledge in host specificity and sexual reproduction between populations found on closely related hosts, resulted in a taxonomy of the plant pathogenic Oomycota with relatively few, but broad taxa (Yerkes and Shaw 1959) that were adopted especially by applied plant pathologists. For Plasmopara halstedii, one of the most relevant sunflower pathogens, no less than ca. 80 genera of the Asteraceae were listed as potential hosts (Leppik 1966) regardless of their evolutionary distance, geographic distribution, and untested susceptibility to strains from different hosts. Agronomic experiences showed that such a species concept was inappropriate and additional classifications on the level of formae speciales were introduced to handle important crop pathogens, as for instance the tobacco blue mould Peronospora hyoscyami de Bary f.sp. tabacina (Adam) Skalický. Meanwhile, molecular tools are available to reinvestigate species delimitation and phylogeny of the Oomycota. In the following chapters we summarize the recent progress in classification and diagnosis of the two major groups of obligate biotrophic (i.e., depending on a living substrate) Oomycota, the downy mildew pathogens, and the white blister rusts. 2.2 Upheaval of Oomycete Taxonomy in the Molecular Era Ten years ago, only very few and punctual genetic data of Oomycota were available in public databases. The species selected for early studies were Achlya bisexualis (Gunderson et al. 1987), Lagenidium giganteum (Förster et al. 1990), and Phytophthora megasperma (Förster et al. 1990); all of these species can be cultured on artificial media. They stood exemplary for the Oomycota and early data served predominantly for supporting the separation of oomycetes from other fungal groups and to confirm the relationships with other straminipilous entities. The sequence loci used for early molecular phylogenetic studies were the mitochondrial cytochrome oxidase (cox2) (Hudspeth et al. 2000) and the nuclear 18S rDNA on the phylogenetic analysis of which Dick et al. (1999) justified the separation of the subclass division of the Peronosporomycetes into Saprolegniomycetidae and Peronosporomycetidae (Fig. 2.1). The molecular era of the obligate biotrophic Oomycota started when Petersen and Rosendahl (2000) sequenced partial 28S rDNA of Peronospora farinosa and Albugo candida to include these organisms in their phylogenetic analysis. Meanwhile, a greater number of genetic loci are accessible for sequence comparison (Hudspeth et al. 2000; Riethmüller et al. 2002; Voglmayr 2003; Thines et al. 2006; Göker et al. 2007) and led to massive taxonomic rearrangements on all levels. The most fundamental steps were perhaps the separation of the white blister rusts from Peronosporomycetidae and the subsequent placement within the new subclass Albuginomycetidae (Thines and Spring 2005) and the reclassification of the graminicolous downy mildews within the Peronosporaceae (Riethmüller et al. 2002; Hudspeth et al. 2003; Thines et al. 2008). Although the former step was not entirely based on molecular data, sequence comparison of nrLSU (Riethmüller 38 O. Spring and M. Thines Oomycetes Peronosporomycetes Lagenidiales Rhipidiomycetidae Rhipidiales Leptomitales Saprolegniomycetidae Leptomitales Salilagenidiales Saprolegniales Sclerosporales Saprolegniales Peronosporales Phytiaceae Albuginaceae Peronosporaceae (Sparrow1960; Dick 1973) 18S rDNA Peronosporomycetidae Pythiales Peronosporales (Dick 2001) Fig. 2.1 Classification of the Oomycota in the pre molecular era and the consequences of first DNA sequences of the nuclear 18S rDNA, before the first comprehensive molecular phylogenies Albuginomycetidae Peronosporomycetes in 2001 and current status of former Peronosporomycetidae Rhipidiomycetidae Saprolegniomycetidae Albuginales Albuginaceae Separation based on phenotypic and molecular data Peronosporomycetidae Pythiales Peronosporales Albuginaceae Albugo Peronospororaceae Basidiophora Benua Bremia Bremiella Paraperonospora Peronospora Plasmopara Pseudoperonospora Molecularbased formation of monophyletic entities Albugo Pustula Wilsoniana Peronosporomycetidae Pythiales Peronosporales Peronospororaceae Basidiophora Benua Bremia Graminivora Hyaloperonospora Novotelnova Paraperonospora Perofascia Peronospora Peronosclerospora Plasmopara Plasmoverna Poakatesthia Protobremia Pseudoperonospora Sclerophthora Sclerospora Viennotia Fig. 2.2 Peronosporomycetes according to Dick (2001) and the current status of the former Peronosporomycetidae. Close relationships to Peronosporaceae of the paraphyletic genus Phytophthora, formerly placed in the Pythiales, have been recognized (Riethmüller et al. 2002; Göker et al. 2007); however, no formal solution for this has yet been proposed et al. 2002), and cox2 (mitochondrial cytochrome C oxidase gene) (Hudspeth et al. 2003; Thines et al. 2008) supported this view, therefore necessisating major changes in the taxonomy of plant parasitic oomycetes (Fig. 2.2). 2 Molecular Techniques for Classification and Diagnosis 39 As a consequence of intensive sampling of molecular genetic data and phylogenetic analysis, eight new genera were described within the Peronosporaceae alone, one genus was relegated into synonymy with Plasmopara, and it was realized that the graminicolous downy mildews are to be placed within the Peronosporaceae. Numerous transfers of taxa were made to newly formed monophyletic entities (Constantinescu and Fatehi 2002; Göker et al. 2004; Voglmayr et al. 2004; Constantinescu et al. 2005; Thines et al. 2006, 2007; Voglmayr and Thines 2007; Voglmayr and Constantinescu 2008). In many cases, this reorganization renewed the search for reliable phenotypic characters which coincide with the molecular genetic classification, a prerequisite for the adoption of such new concepts in applied phytopathology, as was urged previously (Spring and Thines 2004). In particular, it could be shown that fine morphology of the sporangiophores (Constantinescu and Fatehi 2002; Thines 2006), as well as haustoria (Voglmayr et al. 2004; Thines et al. 2006, 2007), and the oospore ornamentation (Voglmayr and Riethmüller 2006; Choi et al. 2007, 2008) are critically important characters. For extensive review see Voglmayr (2008). The ongoing research activities are bidirectional. On the one side, a large number of taxa are still leant at very broad species concepts that require a more precise resolution, especially for large families, such as Fabaceae (Garcia-Blázquez et al. 2008), Brassicaceae (Göker et al. 2004), Lamiaceae, Asteraceae, and Amaranthaceae (Choi et al. 2007). On the other hand, whole genome sequencing has reached Oomycota and will soon accelerate research in physiological and ecological aspects of this group (Birch et al. 2008). The genomes of four Phytophthora species are fully sequenced (for review see Lamour et al. 2007) and selected species of Hyaloperonospora, Albugo, and Pustula are currently under investigation. The results of genome sequencing will be an important source for the search of useful genes for taxonomic and phylogenetic investigations in Oomycota. 2.3 Molecular Tools for Reclassification and Identification of Oomycota The application of molecular tools for the characterization of biotrophic organisms is a particular challenge because the obligate biotrophy hampers cultivation and the accumulation of pathogen material; therefore, these techniques were first used in non-biotrophic groups. Most stages of their life cycle are inevitably and tightly linked to living cells of their hosts. In Oomycetes this is true for all Albuginales (white blister rusts) and for the downy mildews (Peronosporaceae pro partem), whereas members of the second order “Pythiales” – a para- and polyphyletic assemblage – can be cultivated axenically on artificial media (e.g. Pythium, Phytophthora). The only cells of the biotrophic taxa which are accessible without contamination through host material are the spores, and these are often not available in sufficient amounts for several DNA or protein analyzes. For that reason, isozyme analysis and RFLP (restriction fragment length polymorphism) studies which had 40 O. Spring and M. Thines Table 2.1 Molecular-based techniques and their potential use for classification of biotrophic Oomycota as demonstrated for selected groups Research field of Example and reference Useful at taxonomic application level of Isozyme analysis Sub-species Classification Plasmopara (Komjáti et al. 2008) Plasmopara RAPDs Species to sub-species Classification (Roeckel-Drevet et al. 1997) Plasmopara AFLPs Species to sub-species Classification (Roeckel-Drevet et al. 1997) Plasmopara iSSRs Species to sub-species; Classification (Intelmann and Spring 2002; crossing experiments Gobbin et al. 2003) SSU rDNA Kingdom to genus Phylogeny Peronosporomycetes Sequencing (Dick et al. 1999) LSU rDNA Class to species Phylogeny Peronosporomycetes sequencing (Riethmüller et al. 1999, 2002; Petersen and Rosendahl 2000) COX2 sequencing Kingdom to species Phylogeny and Peronosporomycetes (Hudspeth et al. 2000, 2003; calssification Thines et al. 2007; Choi et al. 2007b, 2008) ITS sequencing Family to species Phylogeny and Peronosporoaceae (Göker et al. 2004; Voglmayr classification 2003; Choi et al. 2006, 2007, 2008; Spring et al. 2006; Thines 2007) SNPs Species to sub-species; Phylogeny, Phytophthora; Plasmopara crossing analysis classification, (Martin 2008; Delmotte et al. population 2008) studies – DNA barcoding Species Classification successfully been employed for the investigation of Phytophthora (Oudemans and Coffey 1991; Förster and Coffey 1989) were in general not useful, unless large amounts of sporangia were gained through cultivation of the pathogen on host plants as was recently demonstrated for P. halstedii pathotypes screened by Komjáti et al. (2008). Before PCR (polymerase chain reaction) techniques with oomycete-specific primers had been developed, various so-called fingerprint techniques were used for classification and taxonomic studies (Table 2.1). RAPD (randomly amplified polymorphic DNA) and AFLP (amplified fragment length polymorphism) fingerprints were used for differentiation of isolates of P. halstedii, the downy mildew pathogen of sunflower (Roeckel-Drevet et al. 1997). However, a reliable classification of pathotypes (physiological races) could not be achieved with this technique. Another fingerprint technique is on the basis of the polymorphism of microsatellite markers and simple sequence repeats which are widely distributed elements in eukaryotic genomes (Tautz 1989; Lagercrantz et al. 1993) and revealed to be 2 Molecular Techniques for Classification and Diagnosis 41 helpful for generating highly diverse fingerprint patterns in sporangial DNA samples of some downy mildew pathogens of crop plants (Intelmann and Spring 2002; Gobbin et al. 2003; Spring et al. 2007a; Komjati et al. 2007; Zipper et al. 2009). The amplification patterns allowed the differentiation of two related pathogen species on Xanthium and Helianthus (Komjáti et al. 2007). On the infraspecific level, population studies revealed low genetic variation in the grapevine downy mildew pathogen Plasmopara viticola (Gobbin et al. 2006; Delmotte et al. 2006) and in P. halstedii (Intelmann and Spring 2002), thus prohibiting identification of physiological races. In contrast, differentiation between fungicide sensitive and resistant genotypes of tobacco blue mould Peronospora tabacina based on iSSR polymorphisms was recently shown and used for a population study in field accessions from Europe (Zipper et al. 2009). With the identification of DNA sequences of specific genes, analyzes regarding the identity and evolutionary history of obligate biotrophic oomycetes were greatly advancing. Because of the paucity of material, particularly those genomic regions for which multiple copies per cell exist were chosen for comparative studies. Such prerequisites were found in the repetitive elements of the nuclear ribosomal DNA and in the mitochondrial-encoded cytochrome C oxidase gene (cox2). The small subunit of the 18S rDNA (SSU) was used by Dick et al. (1999) to justify the separation of the Peronosporomycetes into the subclass taxa Saprolegniomycetidae and Peronosporomycetidae while other studies used cox2 for tracing the phylogeny of these entities (Hudspeth et al. 2000, 2003; Cook et al. 2001, Choi et al. 2007, 2008; Thines et al. 2007, 2008). The large subunit of the 28S rDNA (LSU) initially served for the generic resolution of the Saprolegniomycetidae (Riethmüller et al. 1999) and has since been broadly used in many revisions of oomycete taxa (Riethmüller et al. 2002; Göker et al. 2003; Voglmayr et al. 2004; Thines et al. 2006; Voglmayr and Thines 2007). A major advantage of sequence-based classification and identification is its applicability on infected host plant tissue, because specific primers are used for amplifying the target gene in the presence of high amounts of host DNA background. On the other hand, SSU and LSU sequences often do not provide sufficient resolution for the classification of subgeneric taxa, and genomic regions of higher variability were searched for. Meanwhile, the ITS (internal transcribed spacer) region, and to a lesser extent the IGS (inter-genic spacer) region, two noncoding elements of the nuclear rDNA (Bachmann 1994), have become the most preferred sequences for studies on the lower ranking levels (Schurko et al. 2003; Wattier et al. 2003, for additional references see Spring 2004). Thereby, broad species concepts such as the Hyalopernospora complex on Brassicaceae can be tested (Choi et al. 2003; Göker et al. 2003) or the A. candida complex on the same host family (Choi et al. 2006, 2007a, 2008). ITS sequencing is also a useful tool for the classification of pathogens on crop plants, when morphological characters like sporangial size are unsecure. Thus, the downy mildew of poppy (Papaver somniferum) is caused by two phenotypically similar, but not closely related taxa, Peronospora arborescens and P. cristata, and molecular data confirmed the latter to be responsible for epidemics in Tasmania (Scott et al. 2004), while the former is the prominent pathogen of poppy in Europe. 42 O. Spring and M. Thines The potential of variation in ITS has not yet been fully explored. Repetitive elements were recently found in P. halstedii (Thines et al. 2005; Thines 2007) and P. angustiterminalis (Komjáti et al. 2008) which extended this noncoding region to an almost fourfold size in comparison to basal groups of the Peronosporaceae and may enable to trace the speciation process within this lineage. On the infraspecific level, the above-mentioned molecular tools may sometimes provide insufficient resolution. In such cases, SNPs (single nucleotide polymorphisms) provide a promising new type of molecular marker, if sufficient genomic data have been established in the respective group of organisms (Schlotterer 2004). In plant pathogens, SNPs have sparsely been used so far (Morin et al. 2004), and this is true in particular for Oomycota. SNP-based population studies were carried out in Phytophthora ramorum (Martin 2008) and Hyaloperonospora parasitica s.l. (Clewes et al. 2007). In a recent study on P. halstedii field accessions, independent introduction events of the pathogen in French sunflower cultivation were traced on the basis of SNP data (Delmotte et al. 2008). For the classification of unknown samples, genetic barcoding was shown to be a helpful tool (Hebert et al. 2003). In land plant taxonomy, the use of short genetic markers located in mitochondrial or plastidal DNA for the identification of organisms was introduced few years ago (Hebert and Gregory 2005; Chase et al. 2005) and meanwhile standardized protocols have been proposed (Chase et al. 2007). The technique has been adopted for the investigation in bryophytes (Pedersen et al. 2006), algae (Saunders 2008), and partly also for fungi (Seifert et al. 2007). For Oomycota, such attempts have to be considered still preliminary (Göker et al. 2007; Blair et al. 2008), but it is to be expected that a unifying approach will close this gap shortly. The mitochondrial cytochrome C oxydase gene could be one of the candidate genes for genetic barcoding in Oomycota and would allow comparison with other phyla, where this region was also selected (Seifert et al. 2007). Recent advances in the investigation of herbarium specimens (May and Ristaino 2004; Ristaino 2006; Liu et al. 2007; Telle and Thines 2008) are promising to enable the inclusion historical specimens in molecular phylogenetic investigations and molecular barcoding. Telle and Thines (2008) have reported the use of only 2 mg of infected plant tissue of more than 100-year-old specimens. This incites the hope that it might be possible to include type specimens in the investigations, thereby linking the classical, Linnean system to modern molecular barcoding approaches. 2.4 Molecular Approaches for Tracing Occurrence of Oomycetes in Plants and Habitats Molecular techniques provide unprecedented possibilities to identify plant pathogens on or within their hosts and in the environment. In Oomycota, only few reports yet exist on the employment of molecular markers for tracing their occurrence in 2 Molecular Techniques for Classification and Diagnosis 43 host organisms or in soil (e.g. Aegerter et al. 2002; Belbahri et al. 2005; Hukkanen et al. 2006). This is remarkable, taking into account that for many biotrophic species on wild host plants, little more than the mode of sporulation and the disease symptoms are known. Ways of overwintering and developmental stages between penetration of the host and sporulation on its surface are mostly unexplored, even in many economically important diseases of crop plants. Several observations of recurrent and epidemic infections of plants in a certain developmental stage and subsequent seemingly unaffected growth without infection symptoms support the assumption that the pathogen may pass through an asymptomatic or endophytic life stage. Typical examples are the white blister rusts of the genus Albugo on Brassicaceae, for which Jacobson et al. (1998) had postulated an endophytic persistence, because in their study oomycetes were detected by means of specific ITS primers in DNA extracts from symptomless host tissue. According to our own unpublished observations, this appears not to be an exception in biotrophic Oomycota, but may be paralleled by several downy mildew species, e.g. Peronospora veronicae on Veronica spp., and Hyaloperonospora spp. on Cardamine and Erysimum. Such latent types of infection are also known from pathogens on crop plants such as P. halstedii on sunflower (Cohen and Sackston 1974; Spring 2001) or Peronopora sparsa on arctic bramble and boysenberry (Hukkanen et al. 2006). In the latter case, symptomless phases in woody parts and in the root stock of the host (Lindquist et al. 1998) ensure overwintering of the pathogen. Similarly, the survival of the grape downy mildew P. viticola with perennial mycelium in Vitis shoots and dormant buds had been postulated (Pioth 1957; Rumbou and Geissler 2006). The root system and rhizome of perennial hosts can be inhabited by the pathogen that undergoes unrecognized hibernation until symptoms appear on aerial plant parts in the next season. This has been observed for P halstedii in the perennial sunflower Helianthus divaricatus (Nishimura 1922) and more recently in the hybrid species Helianthus x laetiflorus (Spring et al. 2003). In the rootstock of hop plants, hyphae of Pseudoperonospora humuli were detected microscopically, when epidemics of the pathogen had reached for the first time in hop gardens in England (Salmon and Ware 1925). The perennial mycelium is responsible for the occurrence of diseased, stunted shoots in April and May, and gives rise to secondary infection. Tracing pathogen contamination in asymptomatic host tissue is particularly important in plant propagation which is based on cuttings and grafting. Aegerter et al. (2002), for instance, used PCR techniques successfully for the detection of P. sparsa in the shoot cortex and crown tissue of asymptomatic rose plants which were chosen to serve as a source of propagation material. In plant breeding, sensitive tests for the detection of asymptomatic infections could be helpful to avoid false evaluations of seemingly resistant plants. For the quantification of the infection, real-time PCR with downy mildew specific primers was shown to be useful (Hukkanen et al. 2006). Another important developmental stage of symptomless occurrence of vital structures of biothrophic oomycetes involves seeds. Besides ensuring the survival during winter this stage supports the distribution of the pathogen and the spread through seed dispersal. For oomycetes on wild host plants, this phenomenon has not 44 O. Spring and M. Thines been explored yet, whereas seed contamination of crop plants was shown for the downy mildew pathogen of basil (Belbahri et al. 2005) as well as for the white blister rust Pustula on sunflower (Viljoen et al. 1999). With the globalization of seed trading, this way of pathogen distribution has become one of the major problems in agri- and horticulture. Quarantine regulations try to impede seed transmission of diseases, but fast, sensitive and reliable detection methods are still mostly lacking. While germinating and cultivating an appropriate amount of seeds, evaluating the plants for disease symptoms is still the common way of testing seed contamination; molecular techniques have slowly been developed for oomycetes, but tests have not yet been brought into practical usage. For example, sunflower seed contamination with P. halstedii was investigated with various methods. PCR-based detection was attempted with selective oligonucleotide primers (Roeckel-Drevet et al. 1999; Says-Lesage et al. 2000) and with LSU-based primers (Ioos et al. 2007). An ELISA test was developed by Bouterige et al. (2000) and pathogen-specific fatty acids were used for the detection by Spring and Haas (2004). In Peronospora parasitic to basil, specific primers deduced from ITS sequences not only allowed the detection of the downy mildew of basil in seeds and plant tissue, but also enabled quantification of the contamination by using real-time PCR (Belbahri et al. 2005). However, in many cases, the problem for bringing such a test into practice is not only the sensitivity of system, which, in case of sunflower, has proven to be close to market needs and allowed the detection of one contaminated seed out of 50 (Spring and Haas 2004) or even out of 400 (Thines et al. 2004). Perhaps the major obstacle is drawing of a representative sample from large seed batches. A similar problem is the detection of soil contamination with infective structures of the pathogen. Oospores of Oomycota can survive over 10 years or more, hence crop rotation is usually ineffective to prevent yield loss when replanting is made on a previously diseased field. The ability to determine whether a field contains a pathogen is of value both to growers and to researchers, but reports for oomycetes on this topic are still fairly rare. Pratt and Janke (1978) documented the infestation of soil with oospores of Peronosclerospora sorghii and Van der Gaag and Frinking (1997) enriched and extracted oospores of Peronospora viciae from soil by means of sieves. An estimation of inoculum potential for the infection of host plants by planting seeds in oospore contaminated soil was reported for Aphanomyces euteiches, a root pathogen of pea (Malvick et al. 1994) and for the sunflower downy mildew P. halstedii (Gulya 2004). As in the latter case, the applied technique is often a time and resource consuming bioassay that counts the ratio of infected seedlings after planting them in the soil sample under investigation. The adoption of PCR-based techniques to trace oomycetes in soil has so far been limited to Pythium (Wang and Chang 2003) and Phytophthora species (Hussain et al. 2005; Wang et al. 2006; Pavon et al. 2008) and no similar reports exist for downy mildews or white blister rusts. The progress made by molecular approaches in comparison to the bioassay technique is considerable with respect to time consumption and sensitivity. Thus, detection from soil required only 6 h (Wang et al. 2006), whereas bioassays take weeks (Gulya 2004). The limits in sensitivity improved significantly from approximately ten oospores per gram soil (Wang and Chang 2003) to one 2 Molecular Techniques for Classification and Diagnosis 45 oospores per 10 g of soil (Wang et al. 2006), but this depends on the soil screening technique (e.g. Pavon et al. 2008) employed before DNA extraction, and still represents at most a heavily contaminated soil. A burdensome problem for the practical application of such tests is similar as in the seed contamination tests mentioned above, lying in the acquisition of representative soil samples from the fields. The molecular techniques for tracing oomycetes in soil could be used in a modified way also for horticulture where hydroponic irrigation has become popular. Continuous monitoring of the irrigation water could help to identity pathogen contamination at a very early stage and to prevent epidemics in greenhouse cultures. Multiplex detection systems, as recently developed for the detection of fungal and oomycete pathogens of solanaceous crops (Zhang et al. 2008) could help to reduce the costs for the disease management. 2.5 Concluding Remarks The progress made within the past few years in using molecular tools for exploring evolution, taxonomy, and classification of Oomycota is brisk and diminished the distance in knowledge to other eukaryotic organisms rapidly. As the first genomes of plant pathogenic oomycetes were unraveled, the base has been established to resolve the long list of compelling questions. Besides reorganizing the diversity of Oomycota by forming monophyletic entities and splitting unnatural broad taxa, the focus of research will soon shift to aspects of virulence mechanisms, coevolution, oomycete ecology, and agronomically relevant problems. Acknowledgments Funding by the DFG granted to M. Thines is gratefully acknowledged. References Aegerter BJ, Nunez JJ, Davis RM (2002) Detection and management of downy mildew in rose rootstock. Plant Dis 86:1363–1368 Bachmann K (1994) Molecular markers in plant ecology. New Phytol 126:403–418 Bartnicki-Garcia S (1968) Cell wall chemistry, morphogenesis and taxonomy of fungi. Annu Rev Microbiol 22:87–108 Belbahri L, Calmin G, Pawlowski J, Lefort F (2005) Phylogenetic analysis and real time PCR detection of a presumably undescribed Peronospora species on sweet basil and sage. Mycol Res 109:1276–1287 Birch PR, Boevink PC, Gilroy EM, Hein I, Pritchard L, Whisson SC (2008) Oomycete RXLR effectors: delivery, functional redundancy and durable disease resistance. Curr Opin Plant Biol 11:373–379 Bouterige S, Robert R, Marot-Leblond A, Senet J-M (2000) Development of an ELISA test to detect Plasmopara halstedii antigens in seed. In: Proceedings of the 15th International Sunflower Conference, Toulouse, France, pp F44–F49 46 O. Spring and M. Thines Chase MW, Salamin N, Wilkinson M, Dunwell JM, Kesanakurthi RP, Haider N, Savolainen V (2005) Land plants and DNA barcodes: short-term and long-term goals. Philos Trans R Soc B Biol Sci 360:1889–1895 Chase MW, Cowan RS, Hollingsworth PM, van den Berg C, Madrinan S, Petersen G, Seberg O, Jorgensen T, Cameron KM, Carine M, Pedersen N, Hedderson TAJ, Conrad F, Salazar GA, Richardson JE, Hollingsworth ML, Barraclough TG, Kelly L, Wilkinson M (2007) A proposal for a standardised protocol to barcode all land plants. Taxon 56:295–299 Choi Y-J, Hong S-B, Shin H-D (2003) Diversity of the Hyaloperonospora parasitica complex from core brassicaceous hosts based on ITS rDNA sequences. Mycol Res 107:1314–1322 Choi Y-J, Hong S-D, Shin H-D (2006) Genetic diversity within the Albugo candida complex (Peronosporales, Oomycota) inferred from phylogenetic analysis of ITS rDNA and COX2 mtDNA sequences. Mol Phylogenet Evol 40:400–409 Choi Y-J, Shin H-D, Hong S-D, Thines M (2007a) Morphological and molecular discrimination among Albugo candida materials infecting Capsella bursa-pastoris world-wide. Fungal Divers 27:11–34 Choi YJ, Hong SB, Shin HD (2007b) Re-consideration of Peronospora farinosa infecting Spinacia oleracea as distinct species, Peronospora effusa. Mycol Res 110:381–391. Choi Y-J, Shin H-D, Thines M (2008) Evidence for uncharted biodiversity the Albugo candida complex, with the description of a new species. Mycol Res 112:1327–1334 Clewes E, Barbara DJ, Kennedy R (2007) Progress in the development of tools for molecular epidemiology of downy mildew on vegetable in Brassica. In: Lebeda A, Spencer-Phillips PTN (eds) Advances in Downy Mildew Research Vol. 3, Olomouc, Czech Republick, pp 113–118 Cohen Y, Sackston WE (1974) Seed infection and latent infection of sunflower by Plasmopara halstedii. Can J Bot 52:231–236 Constantinescu O, Fatehi J (2002) Peronospora-like fungi (Chromista, Peronosporales) parasitic to Brassicaceae and related hosts. Nova Hedwigia 74:291–338 Constantinescu O, Voglmayr H, Fatehi J, Thines M (2005) Plasmoverna gen. nov., and the taxonomy and nomenclature of Plasmopara (Chromista, Peronosporales). Taxon 54:813–821 Cook KL, Hudspeth DSS, Hudspeth MES (2001) A cox2 phylogeny of representative marine Peronosporomycetes. Nova Hedwigia 122:231–243 Delmotte F, Chen WJ, Richard-Cervera S, Greif C, Papura D, Giresse X, Mondor-Genson G, Corio-Costet MF (2006) Microsatellite DNA markers for Plasmopara viticola, the causal agent of downy mildew of grapes. Mol Ecol Notes 6:379–381 Delmotte F, Giresse X, Richard-Cervera S, M’Baya J, Vear F, Tourvieille J, Walser P, Tourvieille de Labrouhe D (2008) Single nucleotide polymorphisms reveal multiple inroductions into France of Plasmopara halstedii, the plant pathogen causing sunflower downy mildew. Infect Genet Evol 8:534–540 Dick MW (2001) Straminipilous Fungi. Kluwer, Dordrecht, The Netherlands Dick MW, Vick MC, Gibbings JG, Hedderson TA, Lastra CCL (1999) 18S rDNA for species of Leptolegnia and other Peronosporomycetes: justification for the subclass taxa Saprolegniomycetidae and Peronosporomycetidae and division of the Saprolegniacae sensu lato into Leptolegniaceae and Saprolegniaceae. Mycol Res 103:1119–1125 Förster H, Coffey MD (1989) Approaches to the taxonomy of Phytophthora using polymorphisms in mitochondrial and nuclear DNA. In: Lucas JA, Shattock RC, Shaw DS, Cooke LR (eds) Phythophthora. Cambridge University Press, Cambridge, UK, pp 164–183 Förster H, Coffey MD, Elwood H, Sogin ML (1990) Sequence analysis of the small subunit ribosomal RNAs of three zoosporic fungi and implications for fungal evolution. Mycologia 82:306–312 Garcia-Blázquez G, Göker M, Voglmayr H, Martin MP, Telleria MT, Oberwinkler F (2008) Phylogeny of Peronospora, parasitic on Fabaceae, based on ITS sequences. Mycol Res 112:502–512 Gäumann E (1918) Über die Formen der Peronospora parasitica (Pers.) Fries. Beihefte zum Botanischen Zentralblatt 35:395–533 2 Molecular Techniques for Classification and Diagnosis 47 Gäumann E (1923) Beiträge zu einer Monographie der Gattung Peronospora Corda. Beiträge zur Kryptogamenflora der Schweiz 5:1–360 Gobbin D, Pertot I, Gessler C (2003) Identification of microsatellite markers for Plasmopara viticola and establishment of high throughput method for SSR analysis. Eur J Plant Pathol 109:153–164 Gobbin D, Rumbou A, Linde CC, Gessler C (2006) Population genetic structure of Plasmopara viticola after 125 years of colonization in European vineyards. Mol Plant Pathol 7:519–531 Göker M, Voglmayr H, Riethmüller A, Weiß M, Oberwinkler F (2003) Taxonomic aspects of Peronosporaceae inferred from Baysian molecular phylogenetics. Can J Bot 81:1–12 Göker M, Riethmüller A, Voglmayr H, Weiß M, Oberwinkler F (2004) Phylogeny of Hyaloperonospora based on nuclear ribosomal internal transcribed spacer sequences. Mycol Prog 3: 83–176 Göker M, Voglmayr H, Riethmüller A, Oberwinkler F (2007) How do obligate parasites evolve? A multi-gene phylogenetic analysis of downy mildews. Fungal Genet Biol 44:105–122 Gulya TJ (2004) A seedling bioassays to detect the presence of Plasmopara halstedii in soil. In: Spencer-Phillips P, Jeger M (eds) Advances in Downy Mildew Research, Vol 2. Kluwer, Dordrecht, The Netherlands, pp 233–240 Gunderson JH, Elwood H, Ingold A, Kindle A, Sogin ML (1987) Phylogenetic relationships between chlorophytes, chrysophytes and oomycetes. Proc Natl Acad Sci USA 84:5823–5827 Harper JT, Keeling PJ (2003) Nucleus-encoded, plastid-targeted glyceraldehyde-3-phosphate dehydrogenase (GAPDH) indicates a single origin for chromalveolate plastids. Mol Biol Evol 20:1730–1735 Hebert PDN, Gregory TR (2005) The promise of DNA barcoding for taxonomy. Syst Bot 54:852–859 Hebert PDN, Ratnasingham S, DeWaard JR (2003) Barcoding animal life: Cytochrome c oxydase subunit 1 divergences among closely related species. Proc R Soc B Biol Sci 270:96–99 Hudspeth DSS, Nadler SA, Hudspeth MES (2000) A cox2 molecular phylogeny of the Peronosporomycetes. Mycologia 92:674–684 Hudspeth DSS, Stenger D, Hudspeth MES (2003) A cox2 phylogenetic hypothesis for the downy mildews and white rusts. Fungal Divers 13:47–57 Hukkanen A, Pietikäinen L, Kärenlampi S, Kokko H (2006) Quantification of downy mildew (Peronospora sparsa) in Rubus species using real-time PCR. Eur J Plant Pathol 116:225–235 Hussain S, Lees AK, Duncan JM, Cooke DEL (2005) Development of a species-specific and sensitive detection assay for Phytopththora infestans and its application for monitoring of inoculum in tubers and soil. Plant Pathol 54:373–382 Intelmann F, Spring O (2002) Analysis of total DNA by minisatellite and simple-sequence repeat primers for the use of population studies in Plasmopara halstedii. Can J Microbiol 48:555– 559 Ioos R, Laugustin L, Rose S, Tourvieille J, Tourvieille de Labrouhe D (2007) Development of a PCR test to detect the downy mildew causal agent Plasmopara halstedii in sunflower seeds. Plant Pathol 56:209–218 Jacobson DJ, LeFebvre SM, Ojerio RS, Berwald N, Heikkinen E (1998) Persistent, systemic, asymptomatic infections of Albugo candida, an oomycete parasite, detected in three wild crucifer species. Can J Bot 76:739–750 Komjáti H, Walcz I, Viranyi F, Zipper R, Thines M, Spring O (2007) Characteristics of a Plasmopara angustiterminalis isolate from Xanthium strumarium. Eur J Plant Pathol 119:421–428 Komjáti H, Bakony J, Spring O, Viranyi F (2008) Isozyme analysis of Plasmopara halstedii using cellulose acetate gel electrophoresis. Plant Pathol 57:57–63 Lagercrantz U, Ellegren H, Andersson L (1993) The abundance of various polymorphic microsatellite motifs differs between plants and vertebrates. Nucleic Acids Res 21:1111–1115 Lamour K, Win J, Kamoun S (2007) Oomycete genomics: new insights and future directions. FEMS Microbiol Lett 274:1–8 48 O. Spring and M. Thines Leipe DD, Wainright PO, Gunderson JH, Porter D, Patterson DJ, Valois F, Himmerich S, Sogin ML (1994) The stramenopiles from a molecular perspective: 16S-like rRNA sequences from Labyrinthuloides minuta and Cafeteria roenbergensis. Phycologia 33:369–377 Leppik EE (1966) Origin and specialization of Plasmopara halstedii complex on the Compositae. FAO Plant Protect Bull 14:72–76 Lindquist H, Koponen H, Valkonen JPT (1998) Peronospora sparsa on cultivated Rubus articus and its detection by PCR based on ITS sequences. Plant Dis 82:1304–1311 Malvick DK, Paercich JA, Pfleger FL, Givens J, Williams HL (1994) Evaluation of methods for estimating inoculum potential of Aphanomyces euteiches in soil. Plant Dis 78:361–365 Martin FN (2008) Mitochondrial haplotype determination in the oomycete plant pathogen Phytophthora ramorum. Curr Genet 54:23–34 May KJ, Ristaino JB (2004) Identity of the mtDNA haplotype(s) of Phytophthora infestans in historical specimens from the Irish Potato Famine. Mycol Res 108:471–479 Morin PA, Luikart G, Wayne RK (2004) SNPs in ecology, evolution and conservation. Trends Ecol Evol 19:208–216 Nishimura M (1922) Studies in Plasmopara halstedii. J Coll Agric Hokkaido Imperial University 11:185–210 Oudemans P, Coffey MD (1991) Isozyme comparison within and among worldwide sources of three morphologically distinct species of Phytophthora. Mycol Res 95:19–30 Patterson DJ (1989) Stramenopiles: chromophytes from a protistological perspective. In: Green JC, Leadbeater BSC, Diver WL (eds). The chromophyte algae: problems and perspectives. Clarendon Press, Oxford 357–379 Pavon CF, Babadoost M, Lambert KN (2008) Quantification of Phytophthora capsici oospores in soil by sieving-centrifugation and real-time polymerase chain reaction. Plant Dis 92:143–149 Pedersen N, Russell SJ, Newton AE, Ansell SW (2006) A novel molecular protocol for the rapid extraction of DNA from bryophytes and utility of direct amplification of DNA from a single dwarf male. Bryologist 109:257–264 Petersen AB, Rosendahl S (2000) Phylogeny of the Peronosporomycetes (Oomycota) based on partial sequences of the large ribosomal subunit (LSU rDNA). Mycol Res 104:1295–1303 Pioth L (1957) Untersuchungen über anatomische und physiologische Eigenschaften resistenter und anfälliger Reben in Beziehung zur Entwicklung von Plasmopara viticola. Zeitschrift für Pflanzenzüchtung 37:127–158 Pratt RG, Janke GD (1978) Oospores of Sclerospora sorghi in soils of South Texas and their relationships to the incidence of downy mildew in grain sorghum. Phytopathology 68: 1600–1605 Riethmüller A, Weiß M, Oberwinkler F (1999) Phylogenetic studies of Saprolegniomycetidae and related groups based on nuclear large subunit DNA sequences. Can J Bot 77:1790–1800 Riethmüller A, Voglmayr H, Göker M, Weiß M, Oberwinkler F (2002) Phylogenetic relationships of the downy mildews (Peronosporales) and related groups based on nuclear large subunit ribosomal DNA sequences. Mycologia 94:834–849 Ristaino JB (2006) Tracking the evolutionary history of the potato late blight pathogen with historical collections. Outlooks Pest Manage 17:228–231 Roeckel-Drevet P, Coelho V, Tourvieille J, Nicolas P, Tourvieille De Labrouhe D (1997) Lack of genetic variability in French identified races of Plasmopara halstedii, the cause of downy mildew in sunflower Helianthus annuus. Can J Microbiol 43:260–263 Roeckel-Drevet P, Tourvieille J, Drevet JR, Says-Lesage V, Nicolas P, Tourvieille De Labrouhe D (1999) Development of a polymerase chain reaction diagnostic test for the detection of the biotrophic pathogen Plasmopara halstedii in sunflower Helianthus annuus. Can J Microbiol 45:797–803 Rumbou A, Geissler C (2006) Particular structure of Plasmopara viticola populations evolved under Greek Island conditions. Phytopathology 96:501–509 Salmon ES, Ware WM (1925) On the presence of a perennial mycelium in Pseudoperonospora humuli (Miyabe & Takah.) Wils. Nature 116:134–135 2 Molecular Techniques for Classification and Diagnosis 49 Saunders GW (2008) A DNA barcode examination of the red algal family Dumontiaceae in Canadian waters reveals substantial cryptic species diversity. 1. The foliose Dilsea-Neodilsea compley and Weeksia. Can J Bot 86:773–789 Says-Lesage V, Meliala C, Tourvieille J, Nicolas P, Tourvieille de Labrouhe D, Roeckel-Drevet P (2000) Development of a test to diagnose the presence of sunflower downy mildew (Plasmopara halstedii) in seed samples. In: Proceedings of the 15th International Sunflower Conference, Toulouse, France, pp F38–F43 Schlotterer C (2004) The evolution of molecular markers – just a matter of fashion? Nat Rev Genet 5:63–69 Schurko A, Mendoza L, de Cock AWAM, Klassen GR (2003) Evidence for geographic clusters: Molecular genetic differences among strains of Pythium insidiosum from Asia, Australia and the Americas are explored. Mycologia 95:200–2008 Scott JB, Hay FS, Wilson CR (2004) Phylogenetic analysis of the downy mildew pathogen of oilseed poppy in Tasmania, and its detection by PCR. Mycol Res 108:198–205 Seifert KA, Samson RA, de Waard JR, Houbraken J, Levesque AC, Moncalvo J-M, Louis-Seize G, Hebert DN (2007) Prospects for fungus identification using CO1 DNA barcodes, with Penicillium as a test case. Proc Natl Acad Sci USA 104:3901–3906 Spring O (2001) Nonsystemic infections of sunflower with Plasmopara halstedii and their putative role in the distribution of the pathogen. J Plant Dis Protect 108:329–336 Spring O (2004) Potential and limits fort he use of new characters in the systematics of biotrophic oomycetes. In: Spencer-Phillips P, Jeger M (eds) Advances in Downy Mildew Research, Vol.2, Kluwer Academic Publisher, The Netherlands, pp 211–231 Spring O, Haas K (2004) Eicosapentaenoic acid, a possible marker for downy mildew contamination in sunflower. In: Spencer-Phillips P, Jeger M (eds) Advances in Downy Mildew Research, Vol 2. Kluwer, Dordrecht, The Netherlands, pp 241–247 Spring O, Thines M (2004) On the necessity of new characters for classification and systematics of biotrophic Peronosporomycetes. Planta 219:910–914 Spring O, Voglmayr H, Riethmüller A, Oberwinkler F (2003) Characterization of a Plasmopara isolate from Helianthus x laetiflorus based on cross infection, morphological, fatty acids and molecular phylogenetic data. Mycol Prog 2:163–170 Spring O, Bachofer M, Thines M, Riethmüller A, Göker M, Oberwinkler F (2006) Intraspecific relationship of Plasmopara halstedii isolates differing in pathogenicity and geographic origin based on ITS sequence data. European Journal of Plant Pathology 114:309–315 Spring O, Keil S, Zipper R (2007a) Field monitoring reveals two genotypes of Peronospora tabacina in German tobacco cultures. In: Lebeda A, Spencer-Phillips PTN (eds) Advances in Downy Mildew Research, Proceedings of the 2nd International Downy Mildew Symposium, Vol. 3. Olomouc, Kostelec na Hane, Czech Republic, pp 107–111 Spring O, Hammer TR, Tscheschner H, Zipper R (2007b) Interspecific recombinants from dual infection of sunflower with Plasmopara halstedii and P. angustiterminalis. In: Proceedings of the Jahrestagung der Deutschen Botanischen Gesellschaft, Hamburg, p 70 Tautz D (1989) Hypervariability of simple sequences as a general source for polymoprhic DNA markers. Nucleic Acids Res 17:6463–6467 Telle S, Thines M (2008) Amplification of cox2 ( 620bp) from 2 mg of up to 129 years old herbarium specimens, comparing 19 extraction methods and 15 polymerases. PLoS ONE 3:e3584 Thines M (2006) Evaluation of characters available from herbarium vouchers for the phylogeny of the downy mildew genera (Chromista, Peronosporales), with focus on scanning electron microscopy. Mycotaxon 97:195–218 Thines M (2007) Characterisation and phylogeny of repeated elements giving rise to exceptional length of ITS2 in several downy mildew genera (Peronosporaceae). Fungal Genet Biol 44:199–207 Thines M, Spring O (2005) A revision of Albugo (Chromista, Peronosporomycetes). Mycotaxon 92:443–458 50 O. Spring and M. Thines Thines M, Bachofer M, Zipper R, Spring O (2004) PCR-mediated detection of Plasmopara halstedii in sunflower cultivation. Proceedings of the 54th Deutsche Pflanzenschutztagung. Hamburg, Germany, p 219 Thines M, Komjáti H, Spring O (2005) Exceptional length of ITS in Plasmopara halstedii is due to multiple repetitions in the ITS-2 region. Eur J Plant Pathol 112:395–398 Thines M, Göker M, Spring O, Oberwinkler F (2006) A revision of Bremia graminicola. Mycol Res 110:646–656 Thines M, Göker M, Oberwinkler F, Spring O (2007) A revision of Plasmopara penniseti, with implications for the host range of the downy mildews with pyriform haustoria. Mycol Res 111:1377–1385 Thines M, Göker M, Telle S, Ryley M, Mathur K, Narayana YD, Spring O, Thakur RP (2008) Phylogenetic relationships of graminicolous downy mildews based on cox2 sequence data. Mycol Res 112:345–351 Tommerup IC (1981) Cytology and genetics of downy mildews. In: Spencer DM (ed) The Downy Mildews. Academic, London, UK, pp 121–142 Van der Gaag DJ, Frinking HD (1997) Extraction of oospores of Peronospora viciae from soil. Plant Pathol 46:675–679 Viljoen A, van Wyk PS, Jooste WJ (1999) Occurrence of the white rust pathogen, Albugo tragopogonis, in seed of sunflower. Plant Dis 83:77 Vlk W (1939) Über die Geisselstruktur der Saprolegniaceenschwärmer. Arch Protistenk 92:157–160 Vogel HJ (1960) Two modes of lysine synthesis among lower fungi: evoltionary significance. Biochem Biophys Acta 41:172–173 Voglmayr H (2003) Phylogenetic relationships of Peronospora and related genera based on nuclear ribosomal ITS sequences. Mycol Res 107:1–11 Voglmayr H (2008) Progress and challenges in systematics of downy mildews and white blister rusts: new insights from genes and morphology. Eur J Plant Pathol 122:3–18 Voglmayr H, Constantinescu O (2008) Revision and reclassification of three Plasmopara species based on morphological and molecular phylogenetic data. Mycol Res 112:487–501 Voglmayr H, Riethmüller A (2006) Phylogenetic relationships of Albugo species (white blister rusts) on LSU rDNA sequence and oospore data. Mycol Res 110:75–85 Voglmayr H, Thines M (2007) Phylogenetic relationships and nomenclature of Bremiella sphaerosperma (Chromista, Peronosporales). Mycotaxon 100:11–20 Voglmayr H, Riethmüller A, Göker M, Weiss M, Oberwinkler F (2004) Phylogenetic relationships of Plasmopara, Bremia and other genera of downy mildew pathogens with pyriform haustoria based on Bayesian analysis of partial LSU rDNA sequence data. Mycol Res 108:1011–1024 Wang PH, Chang CW (2003) Detection of the low-germination-rate resting oospores of Pythium myriotylum from soil by PCR. Lett Appl Microbiol 36:157–161 Wang Y, Zhang W, Wang Y, Zheng X (2006) Rapid and sensitive detection of Phytophthora sojae in soil and infected soybeans by species-specific polymerase chain reaction assays. Phytopathology 96:1315–1321 Warner SA, Eierman DF, Sovocool GW, Domnas AJ (1982) Cycloartenol-derived sterol biosynthesis in Peronosporales. Proc Natl Acad Sci USA 79:3769–3772 Wattier RAM, Gathercole LL, Assinder SJ, Gliddon CJ, Deahl KL, Shaw DS, Mills DI (2003) Sequence variation of intergenic mitochondrial DNA spacers (mtDNA-IGS) of Phytophthora infestans (Oomycetes) and related species. Mol Ecol Notes 3:136–138 Yerkes WD, Shaw CG (1959) Taxonomy of Peronospora species on Cruciferae and Chenopodiaceae. Phytopathology 49:499–507 Zhang N, McCarthy ML, Smart CD (2008) A Macroarray system for the detection of fungal and oomycete pathogens of solanaceous crops. Plant Dis 92:953–960 Zipper R, Hammer TR, Spring O (2009) PCR-based monitoring of recent isolates of tobacco blue mold from Europe reveals the presence of two genetically distinct phenotypes differing in fungicide sensitivity. Eur J Plant Pathol 123:367–375 Chapter 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs with a Multiphasic Life Cycle Sigrid Neuhauser, Simon Bulman, and Martin Kirchmair Abstract Plasmodiophorids are an enigmatic group of obligate biotrophic pathogens of higher plants. Together with their sister group phagomyxids, which infect stramenopiles, they form the monophyletic eukaryote clade phytomyxids. They have long been treated as a basal group of fungi, but recent molecular phylogenies point to a close affiliation with the protozoan phylum Cercozoa. The soil-borne and plant-associated nature of plasmodiophorids as well as their multi-stage life cycle with zoosporic, plasmodial, and resting stages has hindered comprehensive research on this group. Plasmodiophorids cannot be cultured without their hosts, and direct observations of any stage of the plasmodiophorid life cycle are difficult and time-consuming. Molecular techniques provide valuable tools for the identification and monitoring of organisms which are difficult to assess with traditional approaches – such as plasmodiophorids. Several different immunological or nucleic acid-based techniques, and more recently genomic and proteomic approaches have been used to investigate plasmodiophorids, their life style, and their interactions with their host plants. Nonetheless, advances in knowledge about plasmodiophorids provided by molecular techniques are mainly restricted to the few economically important species that cause diseases of agricultural crops. Although their taxa may be well described, the available phylogenies of phytomyxids are rather incomplete, as they include only a few selected species. A main reason for this bias is that most specimens deposited in herbaria are too old, soaked in fixatives or otherwise unavailable for DNA analyses. To fully understand this group of protists, more research on “rare”, under-recorded species is needed. S. Neuhauser and M. Kirchmair Institute of Microbiology, Leopold Franzens – University Innsbruck, Technikerstr. 25, 6020 Innsbruck, Austria e-mail: Martin.Kirchmair@uibk.ac.at S. Bulman Plant & Food Research, Private Bag 4704, Christchurch, New Zealand Bio-Protection Research Centre, Lincoln University, P.O. Box 84, 7647 Canterbury, New Zealand Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_3, # Springer-Verlag Berlin Heidelberg 2010 51 52 S. Neuhauser et al. In this review, we discuss the impact of molecular techniques on the detection, monitoring, and characterisation of plasmodiophorids. First, we will briefly introduce plasmodiophorid biology and the taxonomic twists and turns the group has taken to reach its current taxonomic position. Development of methods and experimental progress towards better understanding of plasmodiophorids are then sketched, from classical approaches to the recent “-omics” approaches. We will also discuss future implications of molecular methods, which it is hoped will help to improve knowledge about the role of plasmodiophorids within ecosystems. 3.1 Introduction The phytomyxids (plasmodiophorids and phagomyxids) comprise a monophyletic group of eukaryotes which were originally considered as protists, later as fungi, and are now considered as members of the protist supergroup Rhizaria (see below). Partly to avoid specific taxonomic placement, we use the informal term “plasmodiophorids”, as introduced by Braselton (1995), throughout this review. Plasmodiophorids first came to the attention of scientists and society at the end of the nineteenth century when a severe epidemic of clubroot disease destroyed many of the cabbage crops around St. Petersburg, Russia. The economic loss caused by this disease was tremendous and yet the causative organism or agent was unknown at the time. In 1872, the Russian Gardening Society offered a prize to anyone who could identify the cause of the disease, and who could suggest a control for clubroot. Michail Woronin, a Russian botanist and plant pathologist started his research on the epidemic plague in 1873. A few years later, he described the causative organism – Plasmodiophora brassicae – as novel microorganism to the scientific community (Woronin 1877). Woronin observed plasmodia in diseased root parenchyma cells, the metamorphosis of the plasmodia into spores, as well as the hatching of amoebae out of these spores. His discovery of this new organism stimulated interest in this mysterious group of organisms and many plasmodiophorid plant parasites were described during the next decades (Table 3.1). 3.1.1 What are Plasmodiophorids? Plasmodiophorids are obligate intracellular parasites of green plants. They are characterised by a complex life cycle which will be illustrated by the example of Pl. brassicae, the best-studied plasmodiophorid (Fig. 3.1). The life cycle starts with a primary zoospore that attaches to the wall of a root hair of a cruciferous plant. Soon the flagella are retracted and the zoospore encysts. Aist and Williams (1971) demonstrated clearly and in detail that after the primary zoospore attaches to the cabbage root hair, a projectile-like structure (“Stachel”) is formed within a tubular cavity (the “Rohr”) inside the cyst (terminology according to Keskin and Fuchs 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 53 Table 3.1 Plasmodiophorid species on green plants (no sign), *Oomycetes, +Phaeophyceae, # diatoms, $green algae. y ¼ causing hypertrophies in host plants; n ¼ no hypertrophies visible Genus Species Host species Hypertrophies Additional comments Plasmodiophora brassicae Woronin Crucifers y, roots – halophilae Ferd. & Halophila y, petioles Found only once Winge diplantherae (Ferd. & Diplanthera y, internodes Found only once Winge) Ivimey Cook fici-repentis Andreucci Ficus y, branches Found only once bicaudata Feldmann Zostera y, internodes maritima Feldm.-Maz. Triglochinis y, apex Found only once Tetramyxa parasitica K.I. Goebel Ruppia, y – Zannichellia, Potamogeton rhizophaga Lihnell Juniperus n Found only once triglochinis Molliard Triglochin y no resting spores known elaeagni Y. Yendo & Elaeangus y, roots Found only once K. Takase marina Lipkin & Avidor – 1974 n – Octomyxa achlyae Couch, J. Leitn. Achlya* & Whiffen n – brevilegniae Pend. Brevilegnia*, Geolegnia* Sorosphaera veronicae J. Schröt. Veronica y, shoots – radicalis Ivimey Cook Poa, y, root hairs – & Schwartz Molinea, Catabrosa viticola Kirchm., Neuh. Vitis n – & L. Huber Sorodiscus callitrichis Lagerh. & Callitriche y, shoot – Winge radicicolous Ivimey Gynandropsis y – Cook karlingii Ivimey Cook Chara y, internodes – cokeri Goldie-Sm. Pythium* n – Membranosorus heterantherae Ostenf. & Heteranthera y, roots – H.E. Petersen Spongospora subterranea (Wallr.) Solanaceae y, fine roots – Lagerh. nasturtii M. W. Dick Nasturtium y – campanulae (Ferd. & Campanula y, roots Found only once Winge) Ivimey Cook cotulae Barrett Cotula y – n – Ligniera verrucosa Maire & Veronica other hosts A. Tisson n – junci (Schwartz) Maire Juncus other hosts & A. Tisson pilorum Fron. & Gaillat Poa n probably identical with L. junci (continued) 54 S. Neuhauser et al. Table 3.1 (continued) Genus Species Woronina Polymyxa Phagomyxa Maullinia Host species isoetes Palm betae (Němec) Karling Isoetes Beta hypogeae (Borzı́) Karling plantaginis (Němec) Karling polycystis Cornu pythii Goldie-Sm. glomerata (Cornu) A. Fisch aggregata Zopf Numerous hosts leptolegnia Karling graminis Ledingham betae Keskin algarum Karling chattonii (P.A. Dang.) Karling bellerocheae Schnepf odontellae Kühn, Schnepf & Bulman ectocarpi I. Maier, E.R. Parodi, Westermeier & D.G. Müll. Plantago Saprolegniaceae* Pythium* Vaucheria+ Hypertrophies Additional comments n Found only once n probably identical with L. junci n no resting spores known n no resting spores known n – n – n – Mougeotia$ n Oedogonium$ Different grasses Cenopodiaceae Pylaiella+ Ectocarpus+ – n n n Found only once on both hosts only once – – – Found only once – – Bellerochea# Odontella# n n – – Ectocarpus+ – – 1969). The “Rohr” is evaginated to form a bulbous adhesorium which is attached to the host cell wall. The Stachel passes down the “Rohr”, punctures the host wall and then, within 1 s, an amoeboid infection unit (myxamoeba) is injected into the root hair. The total time from adhesorium formation to host penetration is about 1 min (Aist and Williams 1971). Inside the root hair, the myxamoeba develops into a multinucleate plasmodium which cleaves into a sporangiosorus consisting of numerous zoosporangia. Three to sixteen secondary zoospores hatch from each zoosporangium (Ingram and Tommerup 1972). According to Tommerup and Ingram (1971), two secondary zoospores undergo plasmogamy and the binucleate spore infects the roots of the host, but such a fusion of zoospores has not been confirmed by other authors studying the plasmodiophorid life cycle (LudwigMüller and Schuller 2008). Nevertheless, there is consensus that soon after the infection by secondary zoospores, multinucleate secondary plasmodia are formed in a process accompanied by pronounced cell division and the formation of hypertrophic cells of the host plant. At this time the typical symptoms of clubroot disease become obvious. During further development, the plasmodium cleaves 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 55 Fig. 3.1 Life cycle of Plasmodiophora brassicae: (a) Primary zoospores hatching out of resting spores. (b) Encysted zoospore injects its protoplast into the root hair by the help of a special extrusosome (“Rohr”) and a projectile-like structure (“Stachel”). (c) Primary plasmodia in root hairs. (d) zoosporangia in root hairs. (e) Secondary zoospores infect cabbage root or (f) undergo karyogamy according to Ingram and Tomerup (1971); but plasmogamy could not be confirmed by other authors. (g)–(h) Secondary plasmodia are developed. (i) Formation of resting spores. According to Ingram and Tomerup (1971) karyogamy and meioses (R!) take place prior to resting spore formation. Alternative “microcycles” may occur: Mithen and Margath (1992) concluded that secondary plasmodio may also be developed from primary zoospores. Naiki et al. (1984) demonstrated that secondary zoospores can re-infect root hairs and produce further zoosporangia. (a) and (c)–(i) according to Woronin 1878; (b) according to Aist and Wiliams 1971 into numerous resting spores. These resting spores can survive in soil for at least 7–8 years (Jørstad 1923; Nielsen 1933 cited in Karling 1968). The formation of primary zoospores closes the cycle. Recent studies provide evidence that the life cycle may be more complicated. Naiki et al. (1984) demonstrated that secondary zoospores can re-infect root hairs and produce further zoosporangia. Mithen and Magrath (1992) concluded that secondary zoospores may not be necessary to develop secondary plasmodia. Myxamoeba derived from primary zoospores may migrate from root hairs to cells of the cortical tissue and form secondary plasmodia and galls. This view was supported by observations by Narisawa et al. (1996) who successfully infected the roots with single resting spores and detected plasmodia in the cortical root cells. However, the formation of resting spores and therefore a 56 S. Neuhauser et al. complete life cycle could be initiated only when plants were inoculated with a dikaryotic or two monokaryotic resting spores (Narisawa and Hashiba 1998). Although the life cycle of Pl. brassicae is now relatively well known, there is still a lack of knowledge especially with respect of karyogamy. Moreover, the life cycles of other plasmodiophorids differ in some degree or are not fully known. Braselton (2001) summarised the current knowledge of sexuality in plasmodiophorids as “largely indirect and presumptive”. Seven years later, we know only a little more about this topic. A comprehensive knowledge on the spatiotemporal distribution of the stages of the plasmodiophorids life cycle would provide strategies for identification and detection of these important, plant-associated organisms. 3.1.2 The Plasmodiophorids in the Tree of Life Since the description of the first plasmodiophorid, their taxonomic position remained unresolved for a considerable time. When Woronin established the genus Plasmodiophora, he proposed an affiliation to the protists in the sense of Haeckel and considered them as the simplest group of the Myxomycetes (Woronin 1878). Historically, Myxomycetes and therefore the plasmodiophorids were considered as fungi. In his monograph on the parasitic slime-moulds, Cook (1932) discussed them as “some of the simplest, if not the most primitive, parasitic fungi known”. There was much speculation on the evolution of plasmodiophorids in the pre-molecular era. Mycologists placed the plasmodiophorids in the division Mastigomycota with other flagellate “fungi” (Alexopoulos and Mims 1979). The presence of chitin in the cell wall of resting spores led to the conclusion that plasmodiophorids may be related to Chytridiomycetes (Buczacki 1983). On the basis of the ultrastructure of zoospores, Barr (1992) reverted to the view of Woronin (1878) and classed the plasmodiophorids within the Protozoa. The transitional region of the plasmodiophorid flagellum was found similar to that of “many protists such as the amoeboflagellate Naegleria gruberi”, a species belonging to the Heterolobosea classified among the discicristates (Keeling et al. 2008). CavalierSmith (1993) placed the order Plasmodiophorida in the class Phytomyxea, subphylum Proterozoa, phylum Opalozoa, where he grouped protists with tubular mitochondrial cristae that lack plastids, cortical alveoli, and tubular ciliary hairs. The progress of DNA sequencing has allowed a more profound classification of the plasmodiophorid plant parasites. Molecular phylogenies based on 18S ribosomal rDNA data confirmed that plasmodiophorids were not related to fungi, but no linkage with any other eukaryotic group was found (Castlebury and Dormier 1998; Ward and Adams 1998). Using 18S rDNA data, Cavalier-Smith and Chao (1996/ 1997) concluded that the Plasmodiophorida are most closely allied with a group containing chlorarachneans and sarcomonads and placed them in the class Phytomyxea within a phylum temporarily called Rhizopoda. This phylum was then renamed Cercozoa (Cavalier-Smith 1998, 2000; Cavalier-Smith and Chao 2003). 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 57 Protein sequence data indicate that the closest relatives of Cercozoa are the Foraminifera (Keeling 2001). This relationship was confirmed by Archibald and Keeling (2004) who analysed actin and ubiquitin protein sequences of plasmodiophorids. They found a single amino acid residue insertion at the functionally important processing point between ubiquitin monomers, at the same position in which an otherwise unique insertion exists in the cercozoan and foraminiferan proteins. It was concluded that plasmodiophorids are related to Cercozoa and Foraminifera, although the relationships among these groups remained unresolved. Nikolaev et al. (2004) calculated SSU and actin phylogenies of amoeboid eukaryotes. They found that the Phytomyxea was a sister group to a clade consisting of Phaeodarea, core Cercozoa, and Desmothoracida. In 2005, a revision of the classification of unicellular eukaryotes was suggested (Adl et al. 2005). A hierarchical system without formal rank design (class, order, etc.) was adopted. According to this scheme the Phytomyxea are embedded within the Cercozoa in the super-group Rhizaria. It should be a task of future research to resolve the open questions on the “true” alliance of plasmodiophorids. 3.1.3 Phylogenetic Relationships Within the Plasmodiophorids In his monograph of the “Plasmodiophorales”, Karling (1968) accepted 11 genera and 35 species (plus three varieties which are raised to species level in current publications). Since then, six new species and one new genus of phytomyxids were described (Table 3.1). Only eight of these 44 phytomyxid species were included in phylogenetic DNA analyses: the plasmodiophorids Pl. brassicae, Polymyxa graminis Ledingham, Px. betae Keskin, Spongospora subterranea (Wallr.) Lagerh., Sp. narsturtii M.W. Dick, Sorosphaera veronicae J. Schröt., and the phagomyxids Phagomyxa bellerocheae Schnepf and Ph. odontellae Kühn, Schnepf & Bulman. Archibald and Keeling (2004) included five sequences from a total of three plasmodiophorid species in their phylogenetic study based on actin protein sequences. The plasmodiophorid clade was divided into two sub-clades. One was formed by Sp. subterranea and So. veronicae and a second clade consisted of sequences of two Pl. brassicae isolates. The most comprehensive phylogeny was based on 18S rDNA sequences of eight phytomyxid species (Bulman et al. 2001). In that study, two major clades of phytomyxids were found: the plasmodiophorids parasiting green plants and the phagomyxids parasiting diatoms. The two clades were designated as “orders” according to zoological nomenclature (Phagomyxida, Plasmodiophorida). The phagomyxids consisted of two Phagomyxa species (Ph. bellerocheae, Ph. odontellae) and the plasmodiophorids consisted of Sp. subterranea, Sp. nasturtii (¼ subterranea f. sp. nasturtii), Pl. brassicae, So. veronicae, Px. betae, and Px. graminis. A subclade comprising So. veronicae and the two Polymyxa-species was supported with bootstrap values of 100% in neighbour joining and parsimony analysis. 58 S. Neuhauser et al. Although little is known about the “deep” roots of phytomyxid phylogeny, there are some species whose intraspecific phylogeny has been studied in some detail. Polymyxa species are only distinguishable by their host preferences. To resolve whether Px. betae and Px. graminis are distinct taxonomic units, restriction analysis of the ITS1-5.8S-ITS2 ribosomal rDNA region was applied by Ward et al. (1994). This study confirmed that Px. betae is distinct from Px. graminis. Moreover, the latter species could be divided into two subgroups (ribotypes). A third ribotype of Px. graminis was found on Sorghum plants in India (Ward and Adams 1998). The number of ribotypes increased to six when African and Japanese Polymyxa samples were included in the phylogenetic analyses (Ward et al. 2005b). Legrève et al. (2002) divided Px. graminis into five special forms (formae speciales) analysing ITS sequences of a similar set of strains. Legrève and co-workers (2002) argued that, in addition to sequence data, these special forms can be differentiated by specific combination of host range and temperature requirements. Although the f. speciales of Legrève et al. (2002) correspond to the ribotypes of Ward et al. (2005b; Table 3.2, Fig. 3.2), there is no consensus on the naming of these taxa. For a Table 3.2 Infraspecific taxa in Polymyxa graminis according to Ward et al. (2005a, b) and Legrève et al. (2002). Genbank accession numbers for ITS1-5.8S-ITS2 rDNA sequences used in these studies are given. Sequences used in both studies are printed in bold Polymyxa graminis ribotype Polymyxa graminis formae speciales (Ward et al. 2005) (Legreve et al. 2002) ribotype I Y12824 f. sp. temperata AJ311572, AJ311573, AJ3111574 ribotype II Y12826 f. sp. tepida Y12826 ribotype III AJ311580, Y12825 f. sp. tropicalis AJ311575, AJ311576, AJ311580, Y12825 ribotype IV AJ311577, AJ311579 f. sp. subtropicalis AJ311577, AJ311578, AJ311579 ribotype V AJ010424, AM075820, f. sp. colombiana AJ010424 AM075821, AM075822 ribotype VI AM075823 Fig. 3.2 Cladograms of infraspecific taxa within Polymyxa graminis according to Ward et al. (2005b) and Legrève et al. (2002). Ward et al. (2005b) described ribotypes which correspond to formae speciales suggested by Legrève et al. (2002) 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 59 general acceptance of the “new” infraspecific taxonomic units, more data from different genes and more isolates should be analysed. 3.1.4 Economic Importance of Plasmodiophorids A small number of plasmodiophorids are quite well studied, especially plasmodiophorids which cause plant diseases and/or have economic significance as vectors of plant viruses. Loss caused by clubroot disease (Pl. brassicae; Figs. 3.3 and 3.4) was estimated to be approximately 10% of all cruciferous crops in Australia (Faggian et al. 1999). The hypertrophic roots of infested plants are a sink for sugars and therefore the clubbed roots lead to stunted growth of the plants. As further example of the economic impact of clubroot disease, in Nepal (province of Palung) the economic loss was estimated at US$ 1.4 million in 2004 and 2005 (Timila et al. 2008). Another plasmodiophorid of economic significance is Sp. subterranea, the causal agent of powdery scab of potatoes (Figs. 3.5 and 3.6). Powdery scab lesions are usually small, circular, and uniform in size and are surrounded by a fringe of potato skin when mature. As the tuber skin over the pustules ruptures, a shallow depression filled with a brown, powdery mass of spores and broken-down tissue is exposed. Infected tubers are predisposed to other maladies such as Fusarium dry rot during storage. Moreover, Sp. subterranea can transmit the potato mop-top virus (PMTV). Foliar symptoms of PMTV include yellow rings, V shape markings, and blotches, especially on the lower leaves. Stems can also be stunted, giving a “moptop” effect. Tuber symptoms are called “spraing”, rustbrown discoloration, in the form of arcs or rings and flecks that appears with internal rust-coloured spots on tubers (reviewed by Merz 2008). Rhizomania (lit. “crazy root” or “root madness”) is a serious disease of sugarbeet caused by the beet necrotic yellow vein virus (BNYVV) which is transmitted by the plasmodiophorid Px. betae (reviewed by Varrelmann 2007). Root symptoms include a mass of fine, hairy secondary roots giving the taproot a beard-like appearance. This reduces sugar yield because tonnage or sugar content or both are reduced. Losses can amount to 50–70% of root weight and 2–4% or more of sugar content (EPPO 1997). Px. graminis can cause significant yield reductions in cereal crops. Like Px. betae, the plasmodiophorid itself does not obviously harm its host, but it can acquire and transmit a range of plant viruses (reviewed by Kanyuka et al. 2003). The soil-borne wheat mosaic virus (SBWMV), for example, causes serious diseases in many cereal species, including winter wheat, durum wheat, barley, rye, and triticale (Brakke 1971). Losses of 40–50% occur in infected areas of commercial fields in Florida, USA (Kucharek and Walker 1974). A complete life cycle of Px. graminis has only been observed in monocotyledonous plants, but can be a vector for the Indian Peanut Clump Virus as well (IPCV; Ratna et al. 1991). The dicot “groundnut” serves only as an intermediate host and is not a “natural” reservoir for the plasmodiophorid vector. Peanut clump viruses are among the most damaging soil-borne pathogens of 60 S. Neuhauser et al. groundnut, causing crop losses estimated at over US$38 million per year world-wide (Delfosse et al. 1999). 3.2 3.2.1 Experimental Strategies to Detect and Monitor Plasmodiophorids Direct Observations, Bioassays, and Bait Tests “Easy” direct observations are limited to those plasmodiophorids that cause distinct disease symptoms. Hypertrophies caused by some plasmodiophorid species (Table 3.1) can be easily observed with the naked eye. Examples are the clubbed roots of cruciferous plants (Fig. 3.4) or the shoot galls of Veronica spp. caused by So. veronicae (Fig. 3.7). The lesions on potato tubers caused by Sp. subterranea are also easy to recognise (Figs. 3.5 and 3.6). But not all plasmodiophorids cause hypertrophies, so for those without obvious symptoms on their host plants, more sophisticated monitoring methods are needed. Sporosori of So. viticola exhibit a characteristic green–yellow autofluorescence (Fig. 3.8 and 3.9) at 450–490 nm. This autofluorescence has been used to screen grapevine roots for its presence and to assess the distribution of this plasmodiophorid in commercial vineyards (Huber et al. 2006). Detection methods for identifying plasmodiophorids which do not cause hypertrophies on the host plant or exhibit autofluorescence – like Px. graminis or L. junci – are restricted to microscopical screenings of plant roots. One strategy for the detection of fungal or protozoan pathogens in soils and composts are bioassays or bait tests (Ciafardini and Marotta 1989). These techniques are particularly useful for detecting non-culturable, obligate parasites (Noble and Roberts 2004). Sensitive indicator plants are grown in the test material (e.g. soil), and the presence of the pathogen is indicated by the development of typical disease symptoms on the indicator plants. The benefits of bioassays are that the viability and pathogenicity are indicated as well as the presence of the pathogen. Drawbacks are that bioassays can take several weeks and can be hampered by low inoculum levels or by microbial interactions (Christensen et al. 2001). Nevertheless, bioassays and bait tests are standard procedures in different national and international regulations. For example, the German “Ordinance on Biowaste” (BioAbfV) regulates a testing for Pl. brassicae of biowastes which should be spread on agricultural, silvicultural, or horticultural land. The aim of this regulation is to minimise the risk of spreading viable Pl. brassica resting spores. The method is based on the work of Knoll et al. (1980) and Bruns et al. (1994). To evaluate the efficacy of composting on removing pathogenic propagules, meshbags containing Plasmodiophora root galls mixed with compost are incorporated into biowaste. After composting, the samples are mixed with sterilised sand and peat. Test plants (Brassica juncea variety “Vitasso”) are potted into this mixture, and then incubated for 5 weeks. Affected plants are counted and the root gallings are graded. An infection index can be calculated. 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 61 Figs. 3.3–3.9 Host symptoms of plasmodiophorid infestations. Figs. 3.3–3.4: Clubroot disease on cauliflower (3: healthy roots, 4: clubbed roots) caused by Plasmodiophora brassicae (bars ¼ 1 cm). Figs. 3.5–3.6: Powdery scab of potatoes caused by Spongospora subterranean (5: bar ¼ 1 cm; 6: bar ¼ 5 mm). Fig. 3.7: Shoot gall on Veronica persica induced by Sorosphaera veronicae (bar ¼ 5 mm). Figs. 3.8–3.9: Epifluorescence of sporosori of Sorosphaera viticola in vine roots at 450–490 nm (8: bar ¼ 1 mm; 9: bar ¼ 20 mm) 3.2.2 Molecular Approaches Methods based on molecular components of cells have opened new possibilities in the study of plasmodiophorids. These methods allow the detection of plasmodiophorids in host tissue and in environmental samples such as soil or water, which is of great interest for scientists who want to understand the parasite’s life cycle. The methods also permit risk assessments to be made, which are important for breeders and growers of susceptible crops who face economic losses either from the symptoms caused by plasmodiophorids or the viruses they transmit. Many nucleic acidbased and immunological methods have been developed to facilitate an easy, fast, and highly sensitive detection of plasmodiophorids. These can be used for 62 S. Neuhauser et al. large-scale screenings for infected plants in plantations or for pre-planting tests of soils for the presence of the pathogen. Different molecular approaches to identify plasmodiophorids, including their implementation in the field, will be discussed in the following sections. 3.2.2.1 Antibody-Based Methods Immunological test methods are quick and detect mainly living material, but the laborious development of assays and possible cross-reactions with related species or the host plant are serious drawbacks (Ward et al. 2004a, b). The development of antibodies and sensitive test systems which are ideally applicable in the field requires considerable optimisation. This process is expensive compared with the development of nucleic acid-based methods for the identification of plasmodiophorids. Although dip-stick tests or lateral flow devices could facilitate on-site testing, ELISA (enzyme linked immunosorbent assay) protocols are most commonly used, probably because of their wide application in plant virus detection (Merz et al. 2005): the detection of plasmodiophorid parasites and the viruses they transmit can be done with the same equipment. The obligate biotrophic and endophytic nature of plasmodiophorids makes it very important, but difficult, to avoid cross-reactions, especially with the host plant. Even if protocols for the purification and accumulation of plasmodiophorid resting spores are used, the resulting preparations are never completely free of contaminating plant material or other microorganisms (Wakeham and White 1996; Walsh et al. 1996; Delfosse et al. 2000; Qu and Christ 2006a). Two types of antibodies are used: polyclonal antibodies (a mixture of antibodies; usually obtained from immunised animals) and monoclonal antibodies (a specific antibody against a protein or component of the cell). Polyclonal antibodies have been used in ELISA tests for Sp. subterranea (Harrison et al. 1993; Wallace et al. 1995; Walsh et al. 1996; Merz et al. 2005); Px. betae (Walsh et al. 1996; Mutasa-Göttgens et al. 2000; Kingsnorth et al. 2003a), Px. graminis (Delfosse et al. 2000) and Pl. brassicae (Wakeham and White 1996). All authors reported a (semi-)quantitative detection of resting spores in plant material and soil samples. All antibodies used were reported to be highly specific for the targeted pathogen; only between the two Polymyxa species did some cross-reactions occur (Delfosse et al. 2000). Besides ELISA-tests, polyclonal antibodies for Pl. brassicae were used in western blots, dot blots, dip-sticks, and indirect immunofluorescence microscopy (Arie et al. 1988; Wakeham and White 1996). More specific, but also more laborious, is the production of monoclonal antibodies. An ELISA method based on monoclonal antibodies against a glutathioneS-transferase (GST) specific for Px. betae was established (Mutasa-Göttgens et al. 2000; Kingsnorth et al. 2003a). The GST cDNA was cloned and expressed, then purified and used for antibody production. Monoclonal antibodies were used with varying success for Sp. subterranea: Wallace et al. (1995) derived five different monoclonal antibodies recognising zoospores and plasmodia but not cytosori. They 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 63 observed a cross-hybridisation with Px. graminis, whereas Merz et al. (2005) reported the production of a highly sensitive and specific antibody detecting all stages of the life cycle. In plant pathology, immunological tests are especially popular to detect plant viruses. As some viruses need a plasmodiophorid as vector to infect plants, the presence of the virus gives indirect evidence of the presence of the plasmodioporid. For some of the viruses transmitted by plasmodiophorids, serological detection methods are available. An overview is published at DPVweb (see Description of Plant Viruses; Adams and Antoniw 2006, http://www.dpvweb.net/ and the references therein). 3.2.2.2 Nucleic Acid-Based Methods Nucleic acid-based detection methods were quickly appreciated by plant pathologists dealing with plasmodiophorids (Ward et al. 1994; Buhariwalla et al. 1995). The supplementation or substitution of classical observation methods with PCRbased detection methods for the first time allowed large-scale screening and epidemiological studies (Legrève et al. 2000; Legrève et al. 2003; Legrève et al. 2005; Qu and Christ 2006b). However, the obligate biotrophic lifestyle of the plasmodiophorids remains problematic. Plasmodiophorids cannot grow in the absence of their particular host plants and if it is possible to establish a dual system plant–pathogen, it is virtually impossible to keep it without any environmental contamination. Against this background, it is no surprise that, again, only those few plasmodiophorid species causing plant diseases or transmitting plant viruses have been the focus of DNA detection research. During recent years, emphasis has been placed on the development of highly specific detection methods and speciesspecific primers for the plasmodiophorid crop pathogens (see Table 3.3 and the references therein). Most plasmodiophorid DNA-sequences are available from the ITS1-5.8S-ITS2 region of the rDNA tandem repeat. The majority of the species-specific primers have been designed from the highly variable spacers in this region; as a multi-copy gene, the ribosomal repeat is relatively easy to access (Fig. 3.10, Ward and Adams 1998; Bell et al. 1999; Faggian et al. 1999; Bulman et al. 2001; Wallenhammar and Arwidsson 2001; Down and Clarkson 2002; Legrève et al. 2003; Meunier et al. 2003; van de Graaf et al. 2003; Ward et al. 2004a, b, 2005a, b). There were also attempts to find other, unique plasmodiophorid regions for PCR-based detection. For Pl. brassicae, a single copy gene unique to the pathogen was used (Ito et al. 1997, 1999; Wallenhammar and Arwidsson 2001). The single-copy GST-gene (Mutasa-Göttgens et al. 2000) and another single copy gene (Genbank accession number X83745) were used for the detection of Px. betae (Mutasa et al. 1995; Mutasa et al. 1996; Kingsnorth et al. 2003b). Using the latter gene, a nested PCR (nPCR) method and a single-tube nested PCR (stnPCR) method were developed to identify Px. betae without the risk of a co-amplification of host DNA (Ciafardini and Marotta 1989; Mutasa et al. 1995; Mutasa et al. 1996). The RNA transcript of 64 S. Neuhauser et al. Table 3.3 PCR primers used for the detection of plasmodiophorids. Abbreviations of “Reaction type”: PCR ¼ standard PCR. qPCR ¼ real-time PCR, nPCR ¼ nested PCR, stnPCR ¼ single-tube-nested PCR, mRT-PCR ¼ multiplex reverse transcriptase PCR Taxa Gene Primer Primer sequence (50 –30 ) Reaction Reference type Plasmodiophorids rDNA PNS1 gTT ATC Tgg TTg ATC CTg CC PCR Bulman et al. (2001) Pl. brassicae D85819 PBTZS-2 CCg AgT TCg CgT CAg CgT gA stnPCR Ito et al. (1997) Pl. brassicae D85819 PBTZS-3 CCA CgT CgA TCA CgT TgC AAT stnPCR Ito et al. (1997) Pl. brassicae D85819 PBTZS-4 gCT ggC gTT gAT gTA CTg gAA TT stnPCR Ito et al. (1997) Pl. brassicae D85819 PBAW-10 CCC Cgg ggA TCA CgA TAA ATA ACA nPCR Wallenhammar and Arwidsson (2001) Pl. brassicae D85819 PBAW-11 ggA Agg CCg CCC Agg ACT ACC nPCR Wallenhammar and Arwidsson (2001) Pl. brassicae D85819 PBAW-12 gCC ggC CAg CAT CTC CAT nPCR Wallenhammar and Arwidsson (2001) Pl. brassicae D85819 PBAW-13 CCC CAg ggT TCA CAg CgT TCA A nPCR Wallenhammar and Arwidsson (2001) Pl. brassicae rDNA PbITS1 ACT TgC ATC gAT TAC gTC CC nPCR Faggian et al. (1999) Pl. brassicae rDNA PbITS2 ggC ATT CTC gAg ggT ATC AA nPCR Faggian et al. (1999) Pl. brassicae rDNA PbITS6 CAA CgA gTC AgC TTg AAT gC nPCR Faggian et al. (1999) Pl. brassicae rDNA PbITS7 TgT TTC ggC TAg gAT ggT TC nPCR Faggian et al. (1999) Pl. brassicae rDNA CR2 TAT gCC gCA gCA AAg CTC PCR Bulman et al. (2001) Px. betae GST male gAC ATT gCC gCT CTg ACT T qPCR Kingsnorth et al. (2003a, b) Px. betae GST female AATg AgC TgT TgC CTT ATT TTg gA qPCR Kingsnorth et al. (2003a, b) Px. betae GST probe CAA gCA ggC TCA CgC TgC CAT g qPCR Probe Kingsnorth et al. (2003a, b) Px. betae rDNA 698F CAT gTC ggC AAC CgA AAg T qPCR Ward et al. (2004b) Px. betae rDNA 760R Tgg TTC ggg CgCC CAT qPCR Ward et al. (2004b) Px. betae rDNA 718 T Cgg ATT CTT ggA ACg AAT CCg C qPCR Probe Ward et al. (2004b) Px. betae rDNA BET1 CgA ATC gAC TCT CAT TgT CC PCR Bulman et al. (2001) Px. betae X83745 Pb-5a CAg ggg CAg ACg gAT CgC Ag stnPCR Mutasa et al. (1996) Px. betae X83745 Pb-5b CgT CgA gCg CAg TTC TTg gC stnPCR Mutasa et al. (1996) X83745 Pb-6a AgA TgA ggA TgT CAg TCA gg stnPCR Mutasa et al. (1996) Px. betae Px. betae X83745 Pb-4b CTA TgT ggC AAA CCC AAg stnPCR Mutasa et al. (1996) Px. betae X83745 Pb-3a ACg ATg gAC gAC TAT TgA ggg g nPCR Mutasa et al. (1995) X83745 X83745 X83745 X83745 X83745 rDNA rDNA rDNA rDNA rDNA Pb-3b Pb-N3a/2 Pb-N3b/2 PB4for PB4rev Pxfwd1 Pxfwd2 Pxrev7 PxRealF PxRealR Px. graminis Px. graminis Px. graminis Px. graminis Px. graminis Px. graminis f.sp. temperata Px. graminis f.sp. temperata Px. graminis f.sp. temperata Px. graminis f.sp. temperata Px. graminis f.sp. temperata Px. graminis f.sp. tepida Px. graminis f.sp. tepida Px. graminis f.sp. tepida Px. graminis f.sp. tepida Px. graminis f.sp. tepida Px. graminis ribotype II Px. graminis ribotype II Px. gramins ribotype I Px. gramins ribotype I Polymyxa spp. Polymyxa spp. rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA PxRealP 690F 758R 713 T GRA2 PgtempN-F PgtempN-R Pgtemp-F Pgtemp-R Pgtemp-S PgtepN-F PgtepN-R Pgtep-F Pgtep-R Pgtep-S Pg.F2 Pg.R2 Pg.F1 Pg.R1 ITS2mod mITS5rc gCA gCC TAg TCA CAA ATg gCg Tgg Agg AAA ggg ACT TgT CAg TTg CC CAC ACg CCT gAA ATC ATC TAA C gAT ggC CAA TT CTT ACA C CTg Cgg AAg gAT CAT TAg CgTT ggA Agg ATC ATT AgC gTT gAA T gAg gCA TgC TTC CAg ggC TCT CgT CgC TTC TAC CgA TTg gT CCT TgT TAC gAC TTC TTC TTC CTC TAg T CCg gTg AAC AAT Cg CAg CCC gCA TgC ATC TC Cgg ATT gTC gTT CCA AgA A TCA gCA CgT CCA AAg TCC AT gTT CCA AgA ACC CgA Tgg AC AgC gTT gAA TTTg gTC TTg gT TAg CCA ATT CTC CCg AgT TC ggA gTT gCA TCC CgC ATg CgCC ATg ACg gAT TgT CgT T AgT CAg CAC gTC gC CAA AgT CCA TAg CgT TgA ATg gTT gTT gC TTC gAC TTT AgC CAC CgT TT AAT gTg gAT CgT CTC TgT TgC Tg CAC CT TTT gAT CCA ATT CgT gAA Cgg gAT ggA ACg CCC TCg Tgg Tgg ATg Tgg ATC gTC TCT gTT gCT ggA CCT CAT CTg AgA TCT TgC CAA gT AAC ATg Tgg ATT gTg ggC TAT gTg AAC TCC CAT TCT CCA CAA CgC AA gCT gCg TTC TTC CAT CgT TgT gg CCT ACg gAA ACC TTg TTA Cg Mutasa et al. (1995) Mutasa et al. (1995) Mutasa et al. (1995) Meunier et al. (2003) Meunier et al. (2003) Ward and Adams (1998) Ward and Adams (1998) Ward and Adams (1998) Ward et al. (2005a, b) Ward et al. (2005a, b) qPCR Probe qPCR qPCR qPCR Probe PCR PCR PCR qPCR qPCR qPCR Probe PCR PCR qPCR qPCR qPCR Probe PCR PCR PCR PCR PCR PCR Ward et al. (2005a, b) Ward et al. (2004b) Ward et al. (2004b) Ward et al. (2004b) Bulman et al. (2001) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Vaianopoulos et al. (2005) Ward et al. (2005a, b) Ward et al. (2005a, b) Ward et al. (2005a, b) Ward et al. (2005a, b) Ward and Adams (1998) Ward and Adams (1998) (continued) 65 nPCR nPCR nPCR mRT-PCR mRT-PCR PCR PCR PCR qPCR qPCR 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs Px. betae Px. betae Px. betae Px. betae Px. betae Px. graminis Px. graminis Px. graminis Px. graminis Px. graminis Gene Primer Primer sequence (50 –30 ) Polymyxa spp. Polymyxa spp. Sp. subterranea Sp. subterranea Sp. subterranea rDNA rDNA rDNA rDNA rDNA Psp1 Psp2rev SsTQF1 SsTQR1 SsTQP1 Sp. subterranea Sp. subterranea Sp. subterranea Sp. subterranea Sp. nasturtii Sp. nasturtii Sp. nasturtii So. veronicae So. viticola (Plasmodiophorids?) So. viticola (Plasmodiophorids?) rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA rDNA 66 Table 3.3 (continued) Taxa Legrève et al. (2003) Legrève et al. (2003) van de Graaf et al. (2003) van de Graaf et al. (2003) van de Graaf et al. (2003) Sps1 Sps2 Spo8 Spo9 SSN18 SPO2 WC1 SV1 Psvit F TAg ACg CAg gTC ATC AAC CT Agg gCT CTC gAA AgC gCA A CCg gCA gAC CCA AAA CC Cgg gCg TCA CCC TTC A CAg ACA ATC gCA CCC Agg TTC TCA Tg CCT ggg TgC gAT TgT CTg TT CAC gCC AAT CCT TAg AgA Cg CTg ggT gCg ATT gTC TgT Tg CAC gCC AAT ggT TAg AgA Cg ATT ATC TCC ggA TAg TTC TTg gA Agg CAg ACA gAT TTg ACT CT gCA gAC AgA TTT gAC TCT gg gCC gAC AAT CAC ATT CAA CC ACg CgT TCC AAC TTC TTA gAg ggA Reaction type PCR PCR qPCR qPCR qPCR Probe Reference PCR PCR PCR PCR PCR PCR PCR PCR PCR Bell et al. (1999) Bell et al. (1999) Bulman and Marshall (1998) Bulman and Marshall (1998) Down and Clarkson (2002) Down and Clarkson (2002) Bulman et al. (2001) Bulman et al. (2001) This article Psvit R CAT gCC TCT CTg AgT ATC ggT TTC PCR This article S. Neuhauser et al. 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 67 Fig. 3.10 Diagrammatic illustration of species specific PCR primers and their position on the 18S – ITS1 – 5.8S – ITS2 – 28S ribosomal rDNA region. Primer sequences and the corresponding references can be found in Table 3.3 this gene was used in a multiplex reverse-transcriptase PCR (mRT-PCR) for the simultaneous detection of Px. betae and the viruses it transmits (BNYVV, BSBV, and BVQ, Meunier et al. 2003). An advantage of using RNA as template for PCR is that only active cells are detected and consequently positive results will only be obtained when the pathogen is active in the host plant. But, when resting spores need to be detected, this method has the serious drawback that resting spores contain little RNA. For growers and breeders, it is important to detect soil-inoculum 68 S. Neuhauser et al. prior to planting, so DNA-based methods may be preferred. One disadvantage of nucleic acid-based methods is that processing of samples usually has to be done in the lab with specialised equipment and specially trained staff. For other plant pathogens, on-site PCR methods have been described recently (Tomlinson et al. 2005). Results can be obtained within a few hours in the field, but on-site detection is more difficult. For example, there is the need for electric power supply and a high risk of cross-contamination. Specialist staffs are also required. Therefore, the onsite PCR detection methods are to date not technically mature enough to provide robust high-throughput screenings in the field. The intimate contact of plasmodiophorids with their host plants or with soil matrix is the reason for another problem: the sensitivity for all PCR methods was high for purified zoospore or resting spore preparations, but decreased when plasmodiophorids needed to be detected in plant material or soil samples (Faggian et al. 1999). These problems are not restricted to plasmodiophorids: the biggest difficulty in the application of PCR-based methods in phytopathology is to reproducibly extract high quality DNA from plants and soil (Mumford et al. 2006). The presence of various impurities, co-extracted with soil – such as humic acids, polysaccharides, or metal ions – can hamper PCR (Wilson 1997; Robe et al. 2003). Inhibitory compounds from plant tissues (e.g. polyphenolic substances like tannins or polysaccharides) interfere in the DNA extraction process as well as in later processing steps like PCR. These compounds vary between soil types, within plants, between plant species, and with different (soil) management systems. Hence, it is not surprising that many methods have been published for DNA extraction from soil samples (for reviews see Lakay et al. 2007; Robe et al. 2003; Wilson 1997). Microorganisms are unevenly distributed in soil: they can be bound to soil particles or aggregated around organic matter. Therefore, strong sampling strategies are required to detect soilborne pathogens. However, the production of high quality DNA extracts from plants or soil remains the critical step in nucleicacid-based detection systems for plasmodiophorids. 3.3 Plasmodiophorid Genomics A more complete understanding of the different phases of plasmodiophorid life cycles is of great interest. Much effort has been made to develop DNA-based methods for the detection of plant pathogenic plasmodiophorids, but basic questions about how and which biochemical processes the plasmodiophorid manipulates in the host plant remains unresolved. Even in the “molecular era”, species and genus concepts within the plasmodiophorids are mainly based on the morphology of the sporosori (Figs. 3.11–3.16) and on the host plants (Braselton 2001). Of the different stages of plasmodiophorids, only the resting spores can be used to identify and diagnose a plasmodiophorid in the host plant. Observation of zoosporangia, or even more so, plasmodia and zoospores, is time consuming, demanding, and does not allow discrimination at the species level. Molecular tools are needed to 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 69 Figs. 3.11–3.16 Resting spores of different plasmodiophorids: Fig. 3.11: Plasmodiophora brassicae in roots of Brassica oleacea (cauliflower). Fig. 3.12: Ligniera junci in roothairs of Juncus triglumis. Fig. 3.13: Polymyxa graminis in roots of Poaceae. Fig. 3.14: Sorosphaera veronicae in shoot of Veronica persica. Fig. 3.15: Spongospora subterranean on tubers of Solanum tuberosum (potato). Fig. 3.16: Sorosphaera viticola in roots of Vitis berlandieri  riparia SO4 (bar ¼ 10 mm) determine the species or resolve the processes at the plant-parasite interface. Until the first plant genomes were fully sequenced and annotated, few workers met the challenge of studying the molecular basis of plasmodiophorid infection (Buhariwalla and Mithen 1995; Mutasa et al. 1995; Subr et al. 2002; Brodmann et al. 2002; Graf et al. 2004). The fully sequenced and annotated genome of 70 S. Neuhauser et al. Arabidopsis thaliana, which serves as a host plant for Pl. brassicae, opened new possibilities to study plant/host interactions. In recent years the Arabidopsis/Plasmodiophora pathosystem has been increasingly used to understand processes induced by the pathogen and the subsequent physiological changes in the plant (Winkel-Shirley 2002). A search in the ISI Web of Knowledge database (isiknowledge.com) using the term “Plasmodiophora AND Arabidopsis” produces 60 hits (Accession date: 09.10.2008, 20:35 CET). Only 13 of these works were undertaken between 1992, when the first work was published, and 2000 when the Arabidopsis genome was launched (The Arabidopsis Genome Initiative 2000). Since 2001, 47 works have been published, with 29 of these published from 2006 onwards. Therefore, progress in the genome sequencing of other susceptible plants should help us to understand the mode of interaction of host plants and plasmodiophorids to shed light on their biotic interactions. Transcriptome analysis of Arabidopsis plants suffering from clubroot disease has revealed that more than 1,000 genes are differentially expressed in infected roots (Siemens et al. 2006). The use of Arabidopsis microarrays has demonstrated that genes involved in cell division and expansion, as well as genes associated with plant growth hormones like auxin and cytokinin, are upregulated (Siemens et al. 2006). Interestingly, genes involved in pathogen defence showed either no response to an infection or were downregulated shortly after the infection. This indicates an even closer interplay between plasmodiophorid parasites and their host plant than suspected. The susceptible biotic interaction and the partial resistance of some Arabidopsis varieties to clubroot infection have stimulated the study of differences in resistant and susceptible plants at the molecular level (Jubault et al. 2008) as well. Cao et al. (2008) investigated proteome-level changes in the roots of clubbed roots of Brassica napus. They found 20 proteins to be differentially produced when canola plants were challenged by Pl. brassicae. Results indicate that lignin biosynthesis in the host plant decreased, as did enzymes that are involved in the reactive oxygen species metabolism. This again points toward a very close interplay between plasmodiophorids and their hosts. As found in A. thaliana, proteins involved in plant growth hormone pathways were increasingly produced. Arabidopsis and Pl. brassicae have not only been used to study the changes in the host plant after infection but have also been used to identify genes from the plasmodiophorid (Bulman et al. 2006; Bulman et al. 2007). The obligate nature of the parasite means that a mixture of pathogen and plant is obtained from metabolically active plasmodia. To lessen the number of plant genes, suppression subtractive hybridization between the pathogen and plant RNA was used. The annotated genome of the host plant also allowed the exclusion of host-sequences from the analysis. Seventy six new plasmodiophorid genes were identified (Bulman et al. 2006). In a subsequent work, an intron-rich structure of Pl. brassicae genes was described (Bulman et al. 2007). Almost everything we know about plasmodiophorid plant interactions is compiled from experiments with Pl. brassicae, whereas we know very little about the interactions of plasmodiophorids that do not form galls (e.g. Px. graminis). 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 71 Even less is known about species without economic impact. It will be an interesting task for future research to investigate how these plasmodiophorids influence host metabolism and communicate with the plant. 3.4 How Common are Plasmodiophorids? It is unclear whether plasmodiophorids are genuinely rare, or if their abundance is currently underestimated. As an illustration of this, almost nothing is known about plasmodiophorids infecting plants without agricultural value. This lack of information is reflected in the molecular data available: At NCBI Genbank, there are sequences of only 11 phytomyxid species (3 phagomyxids, 8 plasmodiophorids), amounting to a total of 152 nucleotide sequences (Search term “txid37358”, accession date: 26.9.2008, 19:23 CET). Furthermore, only seven of the sequences from plasmodiophorids infecting green plants are from economically “unimportant” species [So. veronicae (4), Ligniera sp. (2), and So. viticola (1)]. These nucleotide sequences were all derived from one isolate each. Genomic survey sequences (GSS) were deposited only for Pl. brassicae (9) and expressed sequence tags (ESTs) have been deposited only for Pl. brassicae (93) and Px. graminis (1). Sixty protein sequences have been obtained from five species, with the vast majority again from Pl. brassicae (49). Only five sequences derived from environmental clone-libraries are assigned as “plasmodiophorid” although our unpublished sequence comparisons show that there are some unrecognised plasmodiophorids assigned as “uncultured fungus” or “uncultured eukaryote”. Given the progress of environmental screenings using clone-libraries (López-Garcı́a and Moreira 2008), this number is low. The idea that the abundance of plasmodiophorids may be underestimated is encouraged by the results of our sampling of Juncus sp. from five different locations in Austria during summer and autumn 2008. When the root hairs were screened by microscopy, Ligniera junci could be observed easily in four samples. In one Juncus sample, an unintentionally co-sampled Poaceae was found to be infested with Px. graminis. At the same location, So. veronicae was previously found in shoots of Veronica persica (Neuhauser et al. 2005). Another Ligniera-positive Juncus sp. was collected in an area where fungal soil-clone libraries were constructed during a 3 year survey on alterations of soil fungal communities. No plasmodiophorid sequences were detected in that study (Oberkofler 2008) although the primers used have been employed to obtain sequences from plasmodiophorids (Ward et al. 1994). In these examples, the random sampling of potential host plants indicated a high abundance and diversity of plasmodiophorids. It will be highly desirable to develop new PCR primers which preferentially amplify phytomyxids to learn more about these biotrophic parasites. In a preliminary experiment of this kind, we have designed primers for the amplification of partial SSU, ITS 1, and 5.8S rDNA from So. viticola (Table 3.3). These primers were successfully used for direct sequencing of different isolates of L. junci, Px. graminis, and So. veronicae 72 S. Neuhauser et al. from plant material. Those primers have not yet been fully tested, but they seem to be suitable to directly sequence plasmodiophorids from different hosts and give good PCR results when used with soil DNA extracts. PCR primers specific for plasmodiophorids as well as the phagomyxids would allow an active inclusion in analyses of different habitats, which would be a valuable step towards a better understanding of this enigmatic group of eukaryotes. As noted earlier, there are a small but increasing number of unassigned plasmodiophorid sequences appearing in DNA-databases. A coordinated study of these sequences together with specific environmental screening has the potential to reveal new diversity among plasmodiophorids. 3.5 Conclusions and Future Research Although, much recent progress has been made in understanding plasmodiophorids and their interactions with their host plants, they remain a cryptic group of organisms. Few species have been studied in detail, and what is known about these raises more questions than answers. The ecological importance of the plasmodiophorids cannot be evaluated, because their distribution and abundance in non-agricultural areas are not known. Classical screenings are time-consuming and laborious as most plasmodiophorids have a biotrophic interaction with their host plants. It will be important to gain data about the abundance of plasmodiophorids in terrestrial, fresh water, and marine ecosystems. Plasmodiophorids have a large impact on their hosts and may have an important role in shaping certain ecosytems. With the progress of DNA-based taxonomy and detection methods, it will be important to define suitable regions for barcoding in plasmodiophorids. To date mostly rDNA sequence data are available and the 18S region seems to be a promising target. Because of their obligate affiliation with their host plants and to decrease the risk of negative-primer bias, it would be of great use to identify a unique region in the plasmodiophorid genome to minimize the risk of cross-reactions with host plants and other endophytic organisms. First studies on Pl. brassicae genes indicate that there are genes which have no similarity with genes from any other organism (Bulman et al. 2006; Bulman et al. 2007). To define if one of these regions is suitable for barcoding of phytomyxids will be task of future research. DNA data from several regions of the plasmodiophorid genomes would also be desirable to create a multigene phylogenetic tree. This would facilitate a better understanding of the intra- and infragenetic relationships between plasmodiophorid species and related organisms and help the plasmodiophorids to occupy a well defined place in the tree of life. As the life cycle of many plasmodiophorids is not resolved in detail, molecular data should be supplemented with morphological data. This would allow a new, more comprehensive classification and characterization of the plasmodiophorids. Random samplings indicate that in certain habitats plasmodiophorids occur in high abundance. To evaluate their possible regulatory or selective role in these 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 73 ecosystems, data on their distribution and their interaction with primary and alternative host plants are needed. It is not known if there is a difference in the interaction with the host plant when primary or secondary plasmodia are formed. The primary sporogenic plasmodia can be formed in different intermediate host plants (Karling 1968; Legrève et al. 2000; Legrève et al. 2005; Qu and Christ 2006b) and do not cause any visible symptoms in the host plant. More information about this life cycle stage and comparative analysis with the primary plasmodia could help to understand the compatible interactions leading to hypertrophic growth of the host plant. This information could provide a guide for the breeding of resistant plants. There remains a great deal to be learnt about the biology of this fascinating and increasingly important group of organisms. Acknowledgements The authors wish to thank Ueli Merz and Lars Huber for providing plasmodiophorid samples. We are indebted to Reinhold Pöder for passing on valuable comments and for his kind support with taking the micrographs. The work was partially supported by the FWF (Austrian Science Fund, grant T379-B16, SN). References Adams MJ, Antoniw JF (2006) DPVweb: a comprehensive database of plant and fungal virus genes and genomes. Nucl Acids Res 34:D382–D385 Adl SM, Simpson AGB, Farmer MA, Andersen RA, Anderson OR, Barta JR, Bowser SS, Brugerolle GUY, Fensome RA, Fredericq S, James TY, Karpov S, Kugrens P, Krug J, Lane CE, Lewis LA, Lodge J, Lynn DH, Mann DG, McCourt RM, Mendoza L, Moestrup O, Mozley-Standridge SE, Nerad TA, Shearer CA, Smirnov AV, Spiegel FW, Taylor MFJR (2005) The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. J Eukaryot Microbiol 52:399–451 Aist JR, Williams PH (1971) The cytology and kinetics of cabbage root hair penetration by Plasmodiophora brassicae. Can J Bot 49:2023–2034 Alexopoulos CJ, Mims CW (1979) Introducory mycology, 3rd edn. Wiley, New York Archibald JM, Keeling PJ (2004) Actin and ubiquitin protein sequences support a cercozoan/ foraminiferan ancestry for the plasmodiophorid plant pathogens. J Eukaryot Microbiol 51:113–118 Arie T, Namba S, Yamashita S, Doi Y (1988) Detection of resting spores of Plasmodiophora brassicae Woron. from soil and root by fluorescent-antibody technique. Ann Phytopathol Soc Jap 54:242–245 Barr DJS (1992) Evolution and kingdoms of organisms from the perspective of a mycologist. Mycologia 84:1–11 Bell KS, Roberts J, Verrall S, Cullen DW, Williams NA, Harrison JG, Toth IK, Cooke DEL, Duncan JM, Claxton JR (1999) Detection and quantification of Spongospora subterranea f. sp. subterranea in soils and on tubers using specific PCR primers. Eur J Plant Pathol 105:905–915 Brakke MK (1971) Soil-borne wheat mosaic virus. Commonwealth Mycological Institute/Association of Applied Biologists Descriptions of Plant Viruses, no 77 Braselton JB (1995) Current status of the plasmodiophorids. Crit Rev Microbiol 21:263–275 Braselton JB (2001) Plasmodiophoramycota. In: Mc Laughlin DJ, Mc Laughlin EG, Lemke PA (eds) The Mycota VII Part A. Springer, Berlin, Heidelberg, pp 81–91 74 S. Neuhauser et al. Brodmann D, Schuller A, Ludwig-Muller J, Aeschbacher RA, Wiemken A, Boller T, Wingler A (2002) Induction of trehalase in Arabidopsis plants infected with the trehalose-producing pathogen Plasmodiophora brassicae. Mol Plant Microbe 15:693–700 Bruns C, Gottschall R, Marchiniszyn E, Schüler C, Zeller W, Wolf G, Vogtmann H (1994) Phytohygiene der Kompostierung – Sachstand, Prüfmethoden, F. und E. Vorhaben. In: Tagungsband BMFT-Statusseminar Neue Techniken der Kompostierung. Hamburg, pp 191–206 Buczacki ST (1983) Plasmodiophora, an interrelationship between biological and practical problems. In: Buczacki ST (ed) Zoosporic plant pathogens, a modern perspective. Academic Press, New York, pp 339–353 Buhariwalla H, Greaves S, Magrath R, Mithen R (1995) Development of specific PCR primers for the amplification of polymorphic DNA from the obligate root pathogen Plasmodiophora brassicae. Physiol Molecul Plant Pathol 47:83–94 Bulman SR, Marshall JW (1998) Detection of Spongospora subterranea in potato tuber lesions using the polymerase chain reaction (PCR). Plant Pathol 47:759–766 Buhariwalla H, Mithen R (1995) Cloning of a Brassica repetitive DNA element from resting spores of Plasmodiophora brassicae. Physiol Mol Plant Pathol 47:95–101 Bulman SR, Kuhn SF, Marshall JW, Schnepf E (2001) A phylogenetic analysis of the SSU rRNA from members of the Plasmodiophorida and Phagomyxida. Protist 152:43–51 Bulman S, Siemens J, Ridgway HJ, Eady C, Conner AJ (2006) Identification of genes from the obligate intracellular plant pathogen, Plasmodiophora brassicae. FEMS Microbiol Lett 264:198–204 Bulman SR, Ridgway HJ, Eady C, Conner AJ (2007) Intron-rich gene structure in the intracellular plant parasite Plasmodiophora brassicae. Protist 158:423–433 Cao T, Srivastava S, Rahman MH, Kav NNV, Hotte N, Deyholos MK, Strelkov SE (2008) Proteome-level changes in the roots of Brassica napus as a result of Plasmodiophora brassicae infection. Plant Sci 174:97–115 Castlebury LA, Dormier LL (1998) Small subunit ribosomal RNA gene phylogeny of Plasmodiophora brassicae. Mycologia 90:102–107 Cavalier-Smith T (1993) The protozoa phylum Opalozoa. J Euk Microbiol 40:609–615 Cavalier-Smith T (1998) Neomonada and the Origin of Animals and Fungi. In: Coombs GH, Vickerman K, Sleigh MA, Warren A (eds) Evolutionary Relationships Among Protozoa. Kluwer, London, pp 375–407 Cavalier-Smith T (2000) Flagellate Megaevolution: the Basis for Eukaryote Diversification. In: Green JR, Leadbeater BSC (eds) The Flagellates. Taylor and Francis, London, pp 361–390 Cavalier-Smith T, Chao EE-Y (1996/1997) Sarcomonad ribosomal RNA sequences, rhizopod phylogeny, and the origin of euglyphid amoebae. Arch Protistenkd 147: 227–236 Cavalier-Smith T, Chao EE-Y (2003) Phylogeny and Classification of Phylum Cercozoa (Protozoa). Protist 154:341–358 Christensen KK, Kron E, Carlsbaek M (2001) Development of a Nordic System for Evaluating the Sanitary Quality of Compost. Nordic Council of Ministers, Copenhagen Ciafardini G, Marotta B (1989) Use of the Most-Probable-Number technique to detect Polymyxa betae (Plasmodiophoromycetes) in soil. Appl Environ Microbiol 55:1273–1278 Cook WRI (1932) The Parasitic slime moulds. Hong Kong Nat Suppl 1:29–39 Delfosse P, Barghava AK, Sobti AK, Reddy AS, Abdurahman MD, Legrève A, Maraite H, Reddy DVR (1999) Geographic distribution of the Indian peanut clump virus (IPCV) in Rajasthan: Soil characteristics and farming practices influencing the disease occurrence. In: Proceedings of the 4th symposium of the international working group on plant viruses with fungal vectors. American Society of sugar beet Technologists, Denver, USA, pp. 109–112 Delfosse P, Reddy AS, Legerve A, Devi KT, Abdurahman MD, Maraite H, Reddy DVR (2000) Serological methods for detection of Polymyxa graminis, an obligate root parasite and vector of plant viruses. Phytopathology 90:537–545 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 75 Down GJ, Clarkson JM (2002) Development of a PCR-based diagnostic test for Spongospora subterranea f.sp. nasturtii, the causal agent of crook root of watercress (Rorippa nasturtiumaquaticum). Plant Pathol 51:275–280 EPPO (1997) Data Sheets on Quarantine Pests. Beet necrotic yellow vein furovirus. In: Smith IM, McNamara DG, Scott PR, Holderness M (eds) Quarantine Pests for Europe, 2nd edn. CABI International, Wallingford Faggian R, Lawrie AC, Bulman SR, Porter IJ (1999) Specific polymerase chain reaction primers for the detection of Plasmodiophora brassicae in soil and water. Phytopathology 89:392–397 Graf H, Fahling M, Siemens J (2004) Chromosome polymorphism of the obligate biotrophic parasite Plasmodiophora brassicae. J Phytopathol 152:86–91 Harrison JG, Rees EA, Barker H, Lowe R (1993) Detection of spore balls of Spongospora subterranea on potato tubers by enzyme-linked-immunosorbent-assay. Plant Pathol 42: 181–186 Huber L, Scholz C, Eisenbeis G, Rühl EH, Neuhauser S, Kirchmair M (2006) Field distribution of Sorosphaera viticola in commercial vineyards in Germany. FEMS Microbiol Lett 260:63–68 Ingram DS, Tommerup IC (1972) The life history of Plasmodiophora brassicae Woron. Proc R Soc Lond B 180:103–112 Ito S, Maehara T, Tanaka S, Kameya Iwaki M, Yano S, Kishi F (1997) Cloning of a single-copy DNA sequence unique to Plasmodiophora brassicae. Physiol Mol Plant Pathol 50:289–300 Ito S, Maehara T, Maruno E, Tanaka S, Kameya-Iwaki M, Kishi F (1999) Development of a PCRbased assay for the detection of Plasmodiophora brassicae in soil. J Phytopathol 147:83–88 Jørstad I (1923) Hvorledes skal man bekaempe klumproten paa vore Kaalvekster? Norsk Havetid 39:126–127 Jubault M, Lariagon C, Simon M, Delourme R, Manzanares-Dauleux MJ (2008) Identification of quantitative trait loci controlling partial clubroot resistance in new mapping populations of Arabidopsis thaliana. Theor Appl Genet 117:191–202 Kanyuka K, Ward E, Adams MJ (2003) Polymyxa graminis and the cereal viruses it transmits: a research challenge. Mol Plant Pathol 4:393–406 Karling JS (1968) The Plasmodiophorales, 2nd completely revised Edition. Hafner Publishing Company, New York, p 256 Keeling PJ (2001) Foraminifera and Cercozoa are related in actin phylogeny: Two orphans find a home? Mol Biol Evol 18:1551–1557 Keeling P, Leander BS, Simpson A (2008) Eukaryotes. Eukaryota, Organisms with nucleated cells. In: The Tree of Life Web Project, http://tolweb.org/ Version 02 September 2008 (under construction). http://tolweb.org/Eukaryotes/3/2008.09.02 Keskin B, Fuchs WH (1969) Der Infektionsvorgang bei Polymyxa betae. Arch Mikrobiol 68:218– 226 Kingsnorth CS, Asher MJC, Chwarszczynska DM, Luterbacher MC, Mutasa-Göttgens ES, Keane GJP (2003a) Development of a recombinant antibody ELISA test for the detection of Polymyxa betae and its use in resistance screening. Plant Pathol 52:673–680 Kingsnorth CS, Chwarszczynska DM, Asher MJC, Kingsnorth AJ, Lyons PA (2003b) Real-time analysis of Polymyxa betae GST expression in infected sugar beet. Mol Plant Pathol 4:171–176 Knoll K-H, Strauch D, Holst H (1980) Standardisierung von Hygieneuntersuchungen für Kompostierungsverfahren. Forschungsbericht 79-10302403, Umweltforschungsplan des MBI, Abfallwirtschaft Kucharek TA, Walker JH (1974) The presence of and damage caused by soilborne wheat mosaic virus in Florida. Plant Dis Rep 58:763–765 Lakay FM, Botha A, Prior BA (2007) Comparative analysis of environmental DNA extraction and purification methods from different humic acid-rich soils. J Appl Microbiol 102:265–273 Legrève A, Vanpee B, Delfosse P, Maraite H (2000) Host range of tropical and sub-tropical isolates of Polymyxa graminis. Eur. J Plant Pathol 106:379–389 76 S. Neuhauser et al. Legrève A, Delfosse P, Maraite H (2002) Phylogenetic analysis of Polymyxa species based on nuclear 5.8 S and internal transcribed spacers ribosomal DNA sequences. Mycol Res 106:138–147 Legrève A, Delfosse P, Van Hese V, Bragard C, Maraite H (2003) Broad- spectrum detection of Polymyxa species and form species by polymerase chain reaction. In: Rush C, Merz U (ed) Proceedings of the fifth Symposium of the International Working Group on Plant Viruses with Fungal Vectors. American Society of Sugar Beet Technologists, Denver Zurich, pp 40–43 Legrève A, Schmit J, Bragard C, Maraite H (2005) The role of climate and alternative hosts in the epidemiology of rhizomanina. In: Rush C (ed) Proceedings of the sixth symposium of the international working group on plant viruses with fungal vectors. American Society of Sugar Beet Technologists, Denver Zurich, pp 125–129 López-Garcı́a P, Moreira D (2008) Tracking microbial biodiversity through molecular and genomic ecology. Res Microbiol 159:67–73 Ludwig-Müller J, Schuller A (2008) What can we learn from clubroots: alterations in host roots and hormone homeostasis caused by Plasmodiophora brassicae. Eur J Plant Pathol 121:291–302 Merz U, Walsh JA, Bouchek-Mechiche K, Oberhänsli T, Bitterlin W (2005) Improved immunological detection of Spongospora subterranea. Eur J Plant Pathol 111:371–379 Merz U (2008) Powdery scab of potato – Occurrence, life cycle and epidemiology. Am J Potato Res 4:241–246 Meunier A, Schmit J-F, Stas A, Bragard C, Kutluk N (2003) Multiplex reverse transcription-PCR for simultaneous detection of Beet necrotic yellow vein virus, Beet soilborne virus, and Beet virus Q and their vector Polymyxa betae Keskin on sugar beet. Appl Environ Microbiol 69:2356–2360 Mithen R, Magrath R (1992) A contribution to the life history of Plasmodiophora brassicae: secondary plasmodia develop in root galls of Arabidopsis thaliana. Mycol Res 96:877–885 Mumford R, Boonham N, Tomlinson J, Barker I (2006) Advances in molecular phytodiagnostics – New solutions for old problems. Eur J Plant Pathol 116:1–19 Mutasa ES, Chwarszczynska DM, Adams MJ, Ward E, Asher MJC (1995) Development of PCR for the detection of Polymyxa betae in sugar beet roots and its application in field studies. Physiol Mol Plant Pathol 47:303–313 Mutasa ES, Chwarszczynska DM, Asher MJC (1996) Single-tube nested PCR for the diagnosis of Polymyxa betae infection in sugar beet roots and colorimetric analysis of amplified products. Phytopathology 86:493–497 Mutasa-Göttgens ES, Chwarszczynska DM, Halsey K, Asher MJC (2000) Specific polyclonal antibodies for the obligate plant parasite Polymyxa – a targeted recombinant DNA approach. Plant Pathol 49:276–287 Naiki T, Kawaguchi C, Ikegami H (1984) Root hair reinfection in Chinese cabbage seedlings by secondary zoospores of Plasmodiophora brassicae Woronin. Ann Phytopathol Soc Japan 50:216–220 Narisawa K, Kageyama K, Hashiba T (1996) Efficient root infection with single resting spores of Plasmodiophora brassicae. Mycol Res 100:855–858 Narisawa K, Hashiba T (1998) Development of resting spores on plants inoculated with a dikaryotic resting spore of Plasmodiophora brassicae. Mycol Res 102:949–952 Neuhauser S, Huber L, Kirchmair M (2005) Sorosphaera veronicae; neu für Österreich. Österreichische Zeitschrift für Pilzkunde 14:303–308 Nielsen NJ (1933) Forsøg med Bekaempelse af Kaalbroksvamp. Tidsskr Planteavl 39:361–391 Nikolaev SI, Berney C, Fahrni JF, Bolivar I, Polet S, Mylnikov AP, Aleshin VV, Petrov NB, Pawlowski J (2004) The twilight of Heliozoa and rise of Rhizaria, an emerging supergroup of amoeboid eukaryotes. Proc Natl Acad Sci USA 101:8066–8071 Noble R, Roberts SR (2004) Eradication of plant pathogens and nematodes during composting: a review. Plant Pathology 53:548–568 Oberkofler I (2008) Soil fungal communities in an alpine primary successional habitat. PhD thesis, Leopold Franzens University Innsbruck 3 Plasmodiophorids: The Challenge to Understand Soil-Borne, Obligate Biotrophs 77 Qu XS, Christ BJ (2006a) Single cystosorus isolate production and restriction fragment length polymorphism characterization of the obligate biotroph Spongospora subterranea f. sp subterranea. Phytopathology 96:1157–1163 Qu XS, Christ BJ (2006b) The host range of Spongospora subterranea f. sp subterranea in the United States. Am J Potato Res 83:343–347 Ratna AS, Rao AS, Reddy AS, Nolt BL, Reddy DVR, Vijayalakshmi M, McDonald D (1991) Studies on transmission of Indian peanut clump virus disease by Polymyxa graminis. Ann Appl Biol 118:71–78 Robe P, Nalin R, Capellano C, Vogel TM, Simonet P (2003) Extraction of DNA from soil. Eur J Soil Biol 39:183–190 Siemens J, Keller I, Sarx J, Kunz S, Schuller A, Nagel W, Schmulling T, Parniske M, LudwigMuller J (2006) Transcriptome analysis of Arabidopsis clubroots indicate a key role for cytokinins in disease development. Mol Plant Microbe 19:480–494 Subr ZW, Kastirr U, Kühne T (2002) Subtractive cloning of DNA from Polymyxa graminis – An obligate parasitic plasmodiophorid. J Phytopathol 150:564–568 The Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796–815 Timila RD, Correll JC, Duwadi VR (2008) Severe and widespread clubroot epidemics in Nepal. Plant Dis 96:317 Tomlinson JA, Boonham N, Hughes KJD, Griffen RL, Barker I (2005) On-site DNA extraction and real-time PCR for detection of Phytophthora ramorum in the field. Appl Environ Microbiol 71:6702–6710 Tommerup IC, Ingram DS (1971) The life-cycle of Plasmodiophora brassicae Woron. in Brassica tissue cultures and in intact roots. New Phytol 70:327–332 Vaı̈anopoulos C, Bragard C, Maraite H, Legréve A (2005) Methods for detection of Polymyxa graminis f. sp. temperata and Polymyxa graminis f. sp. tepida in association with bymo- and furoviruses. Proceedings of the 6th Symposium of the International Working Group on Plant Viruses with Fungal Vectors (IWGPVFV), Bologna, Italy, p 18 van de Graaf P, Lees AK, Cullen DW, Duncan JM (2003) Detection and quantification of Spongospora subterranea in soil, water and plant tissue samples using real-time PCR. Eur J Plant Pathol 109:589–597 Varrelmann M (2007) Occurrence, spread and pathogenicity of different forms of the Rhizomania virus (Beet necrotic yellow vein virus, BNYVV) – Review on biology and variability of Rhizomania and on detection of isolates possibly overcoming resistance. Zuckerindustrie 132:113–120 Wakeham AJ, White JG (1996) Serological detection in soil of Plasmodiophora brassicae resting spores. Physiol Mol Plant Pathol 48:289–303 Wallace A, Williams NA, Lowe R, Harrison JG (1995) Detection of Spongospora-subterranea using monoclonal-antibodies in ELISA. Plant Pathol 44:355–365 Wallenhammar A-C, Arwidsson O (2001) Detection of Plasmodiophora brassicae by PCR in naturally infested soils. Eur J Plant Pathol 107:313–321 Walsh JA, Merz U, Harrison JG (1996) Serological detection of spore balls of Spongospora subterranea and quantification in soil. Plant Pathol 45:884–895 Ward E, Adams MJ, Mutasa CR, Collier CR, Asher MJC (1994) Characterisation of Polymyxa species by restriction analysis of PCR-amplified ribosomal DNA. Plant Pathol 43:872–877 Ward E, Adams M (1998) Analysis of ribosomal DNA sequences of Polymyxa species and related fungi and the development of genus- and species-specific primers. Mycol Res 102:965–974 Ward E, Foster SJ, Fraaije BA, McCartney AH (2004a) Plant pathogen diagnostics: immunological and nucleic acid-based approaches. Ann Appl Biol 145:1–16 Ward E, Kanyuka K, Motteram J, Kornyukhin D, Adams MJ (2005a) The use of conventional and quantitative real-time PCR assays for Polymyxa graminis to examine host plant resistance, inoculum levels and intraspecific variation. New Phytol 165:875–885 Ward E, Motterham J, Kanyuka K, Adams MJ (2005b) The use of PCR methods for Polymyxa graminis to study intraspecific variation, phylogeny and inoculum levels. In: Rush, C (ed) 78 S. Neuhauser et al. Proceedings of the 6th symposium of the international working group on plant viruses with fungal vectors. American Society of sugar beet Technologists, Denver, pp 100–103 Ward LI, Fenn MGE, Henry CM (2004b) A rapid method for direct detection of Polymyxa DNA in soil. Plant Pathol 53:485–490 Wilson IG (1997) Inhibition and facilitation of nucleic acid amplification. Appl Environ Microbiol 63:3741–3751 Winkel-Shirley B (2002) Biosynthesis of flavonoids and effects of stress. Curr Opin Plant Biol 5:218–223 Woronin M (1877) Trudy St Petersb Obshch Est Otd Bot 8:169 Woronin M (1878) Plasmodiophora brassicae. Urheber der Kohlpflanzen Hernie. Jahrbuch für wissenschaftliche Botanik 11:548–574 Chapter 4 Applications of Molecular Markers and DNA Sequences in Identifying Fungal Pathogens of Cool Season Grain Legumes Evans N. Njambere, Renuka N. Attanayake, and Weidong Chen Abstract Molecular techniques have now been widely applied in many disciplines of biological sciences including fungal identification in microbial ecology and in plant pathology. In plant pathology, it is now common to use molecular techniques to identify and study plant pathogens of many agronomical and horticultural crops including cool season grain legume crops. In this chapter, we present two examples in which molecular techniques have been applied in order to identify and investigate multiple fungal pathogens causing two important diseases of chickpea and lentil. In each case, molecular techniques improved over traditional morphological identification and allowed timely and unambiguous identification of fungal pathogens. The first example involves identification of two Sclerotinia species (S. sclerotiorum and S. trifoliorum) causing stem rot of chickpea. Traditional method requires induction of carpogenic germination and observation of dimorphic ascospores in S. trifoliorum, which takes up to eight weeks. Taking advantage of the group I introns present in the nuclear small subunit rDNA of S. trifoliorum but absent in the same DNA region of S. sclerotiorum, a simple PCR amplification of the targeted DNA region allowed timely and reliable differentiation and identification of the species. The second example is of powdery mildew of lentil. Identification of powdery mildew fungi requires observing the teleomorphic (sexual) state of the pathogens, but this is not always available. In studying lentil powdery mildew in the US Pacific Northwest, we found that the powdery mildew on lentil does not fit previously reported species (Erysiphe pisi and E. diffusa). Further investigation confirmed that the lentil powdery mildew in the US is E. trifolii, a new pathogen of lentil. This discovery was mainly based on the rDNA ITS sequences and further confirmed by morphological and pathogenicity E.N. Njambere and R.N. Attanayake Department of Plant Pathology, Washington State University, Pullman, WA 99164, USA W. Chen Department of Plant Pathology, Washington State University, Pullman, WA 99164, USA USDA ARS Grain Legume Genetics and Physiology Research Unit, Washington State University, Pullman, WA 99164, USA e-mail: w-chen@wsu.edu Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_4, # Springer-Verlag Berlin Heidelberg 2010 79 80 E.N. Njambere et al. studies. These two examples demonstrate the important role of modern molecular techniques in solving practical agricultural problems. The ITS and adjacent rDNA could be ideal target regions for developing DNA barcodes for identifying these and related fungal species. 4.1 Introduction Cool season grain legumes (chickpea, Cicer arietinum; faba bean, Vicia faba; lentil, Lens culinaris, and pea, Pisum sativum) are important crops worldwide. They are staple food crops in West Asian and North African countries and are important rotational and specialty crops in developed nations. Fungal diseases are important constraints in grain legume productions. Accurate identification of the fungal pathogens is in many cases a prerequisite for effective management of the diseases they cause and for ecological and population genetics studies. However, many fungal species are similar morphologically, and accurate species identification can be difficult. With current advances in biotechnology, molecular genetic markers have been employed for rapid identification of different kinds of fungi (White et al. 1990; Lieckfeldt and Seifert 2000; Njambere et al. 2008; Attanayake et al. 2009). The development of gene-specific primers for PCR amplification (White et al. 1990) has facilitated systematic studies, and the detection and identification of fungal pathogens. The internal transcribed spacer (ITS) region of nuclear ribosomal DNA has generally been considered a convenient marker for molecular identification of fungi at species level because of its conserved feature within species and multi-copy number per genome (Sanchez-Ballesteros et al. 2000). Henry et al. (2000) identified the fungus Aspergillus at species level and differentiated it from other true pathogenic and opportunistic molds using ITS 1 and ITS 2, allowing for early diagnosis and screening of effective antifungal agents for patients. Schneider et al. (1997) developed a method for detection of Rhizoctonia solani isolates, pathogenic and nonpathogenic to tulips, using ITS rDNA sequences, and they could further identify various anastomosis groups. Recent advancement in identifying fungal species using DNA markers is to develop DNA barcodes using species-specific oligonucleotides that are diagnostic of targeted species (Druzhinina et al. 2005). Such specific DNA regions need to be explored for different groups of fungi. In this chapter, we present two examples of applying molecular techniques in identifying fungal pathogens of cool season grain legumes. 4.2 Sclerotinia Stem Rot of Chickpea Sclerotinia stem rot (Fig. 4.1a) is an important disease of chickpea under conducive environmental conditions and is caused by three species of Sclerotinia: S. sclerotiorum, S. minor, and S. trifoliorum (Bretag and Mebalds 1987). Sclerotinia minor 4 Applications of Molecular Markers and DNA Sequences 81 Fig. 4.1 Symptoms and signs of Sclerotinia stem rot of chickpea caused by Sclerotinia trifoliorum (a), and powdery mildew of lentil caused by Erysiphe trifolii (b) can be easily differentiated from the other two species based on its numerous, scattered small-sized sclerotia in culture and in the field. Morphological difference between S. sclerotiorum and S. trifoliorum is subtle. The ultimate differentiation between S. sclerotiorum and S. trifoliorum requires observation of ascospore morphology which entails carpogenic germination of sclerotia. Ascospores of S. trifoliorum show spore-size dimorphism (two different-sized ascospores within a single ascus), whereas ascospores of S. sclerotiorum show no dimorphism (Kohn 1979; Uhm and Fujii 1983a, b). Induction of carpogenic germination of sclerotia of Sclerotinia spp. is a time-consuming process, and may take up to several months. To further complicate the matter, some isolates of S. trifoliorum are heterothallic and require mating with a compatible strain for carpogenic germination and ascospore production (Uhm and Fujii 1983a, b). Even though the process of identifying members of the genus Sclerotinia through sclerotia and other morphological characteristics has been refined over time (Kohn 1979; Rehnstrom and Free 1993), there are limitations to this approach. For instance, the differentiation of S. trifoliorum from S. sclerotiorum based on sclerotial characteristics is difficult because of instability of some sclerotia characteristics with subsequent sub-culturing (Cother 1977). Therefore, to facilitate the separation of the two species, research efforts have been made in searching for molecular techniques that are reliable and convenient to use. Power et al. (2001) reported that S. trifoliorum contains group I introns in the nuclear small subunit rDNA, whereas S. sclerotiorum and S. minor do not contain any introns in the same DNA region. Molecular analysis of the ITS region can eliminate many of the problems associated with the morphological characters and culturing. Analysis of ITS sequence is usually applied to determine species identity (or sometimes higher taxonomic categories) and to identify and discriminate populations within a species. In the genus Sclerotinia, the ITS region is generally not sufficiently variable to distinguish within species diversity; however, the nuclear small subunit rDNA (nSSRrDNA) has been used for this type of study (Holst-Jensen et al. 1999; Power et al. 2001). In this study we explored the differences in the ITS and the nuclear small subunit regions of the rDNA between the two species causing Sclerotinia stem rot of chickpea. 82 4.2.1 E.N. Njambere et al. DNA Isolation and ITS Sequence Analysis DNA was isolated from mycelial mats or sclerotia using the standard extraction procedures such as the FastDNA1 kit described by Chen et al. (1999). DNA quality was checked using agarose gel electrophoresis and quantified using the NanoDropTM spectrophotometer (NanoDrop Technologies, LLC, Wilmington, Delaware, USA) and the concentration adjusted accordingly before PCR amplification. In our study PCR amplifications were conducted using primers ITS1 and ITS4 described by White et al. (1990). The PCR products were verified by agarose gel electrophoresis and purified for direct PCR sequencing using ABI PRISM 377 automatic sequencer (Applied Biosystems, Foster City, CA, USA). Sometimes the PCR fragments are cloned before sequencing. Sequences were determined on both strands for each of the isolates and were aligned for comparison. Most sequence comparisons are carried out using BLASTn (http://www.ncbi.nlm.nih. gov/BLAST) analysis which aligns two or more homologues to detect for presence of one or more ambiguous region within the segments under comparison. Using nine isolates from S. slerotiorum and S. trifoliorum, we amplified a 540 bp DNA fragment of the ITS region (Njambere et al. 2008). Sequence alignment among the nine isolates identified two single nucleotide polymorphic sites (SNPs) within this homologous region that differentiated the two groups of isolates. The two SNPs were located at position 120 (transversion T ! G) and position 376 (transition T ! C) of the amplicon. Three sequences of the isolates were deposited in the GenBank and assigned accession numbers EU082464, EU082465, and EU082466. BLASTn analysis of the ITS locus of some of the chickpea isolates (including EU082464, EU082465) displayed 100% homology to ITS locus of S. trifoliorum in the GenBank, whereas the ITS region of the other isolates (including EU082466) were identical to GenBank S. sclerotiorum isolates. These results therefore suggest that these two SNPs could be used as markers to separate S. sclerotiorum from S. trifoliorum. Although the ITS sequences allowed differentiation between S. sclerotiorum and S. trifoliorum, this technique is not convenient for routine identification because it requires DNA sequencing. 4.2.2 Detection of Group I Introns Group I introns are ribozymes (RNA enzymes) that catalyze chemical reactions, splicing themselves off of their precursors. Group I introns are widely distributed in bacteria, lower eukaryotes, and higher plants. They can be found in genes encoding for rRNA, mRNA, and tRNA, but seem only in rRNA genes in the nuclear genome of lower eukaryotes. No biological functions are known for the group I introns except for splicing themselves off the primary transcripts. Although group I introns are known to spread from location to location and from one organism to another in evolutionary time, they are quite stable and their locations are highly conserved. Thus, if differences in existence of group I introns are found between two species, 4 Applications of Molecular Markers and DNA Sequences 83 the introns provide convenient markers for separation of the species because they can be easily detected through PCR and agarose gel electrophoresis. That is the case for Sclerotinia spp. It was reported by Power et al. (2001) that S. trifoliorum contains group I introns in the nuclear small subunit rDNA, whereas S. sclerotiorum as well as S. minor does not contain group I introns in the same DNA region. We applied this knowledge in identifying S. trifoliorum from twelve isolates collected from chickpea plants. PCR amplifications were done using primer pairs ITS5/ITS4 and NS3/NS6 (White et al. 1990) in an attempt to detect presence or absence of introns in the nuclear small subunit regions of the rDNA. The reaction conditions were identical to those described above for PCR amplification of the ITS region. One or more group I introns were detected in all isolates of S. trifoliorum, and no group I introns were observed at any isolates of S. sclerotiorum (Fig. 4.2). Amplification with PCR can facilitate detection of the group 1 introns using PCR primer flanking the introns. Isolates with introns produce larger PCR products than isolates without introns, which can be easily detected using agarose gel electrophoresis (Fig. 4.2). To be certain that the isolates harboring the group I introns are indeed S. trifoliorum, nine isolates were selected and induced to germinate carpogenically using a method as previously described (Njambere et al. 2008). For the isolates that germinated carpogenically, all the isolates that harbored introns in the rDNA region produced dimorphic ascospores, the ultimate criterion of identifying S. trifoliorum (Fig. 4.3). These confirmatory tests suggest that the group I introns in the rDNA region could be used for a quick and accurate identification of S. trifoliorum at Fig. 4.2 Agarose gels of PCR amplification of the ITS and the nuclear small subunit rDNA regions of Sclerotinia spp. PCR products with primer pairs ITS1 and ITS4 (top) are monomorphic in size (no introns), whereas PCR products with primers ITS4 and ITS5 (bottom) are polymorphic in size (due to presence of introns). The lanes 1, 4, 5, and 11: S. sclerotiorum isolates (without introns); Lanes 2, 3, 6, 7, 8, 9, 10, and 12: S. trifoliorum (with introns) 84 E.N. Njambere et al. Fig. 4.3 Ascospore morphology of Sclerotinia trifolorium (a and b) showing size dimorphism, and of S. sclerotiorum (c) showing no size dimorphism species level. We have employed this technique in identifying more than 100 isolates of S. trifoliorum for population genetic studies. 4.3 Powdery Mildew of Lentil Powdery mildew is a plant disease caused by many different species of fungi in the order Erysiphales (Glawe 2008). The disease occurs in a wide range of plants. Its symptoms are very distinctive, powdery like spots on leaves and stems. The disease can reduce the yield and quality of many crops and commercial values of ornamental plants. In the field crop lentil, it can be a severe disease on certain cultivars and in some parts of the world, particularly in India during January to February (Agrawal and Prasad 1997). Although lentil is a field crop, breeding materials and many experimental plants are produced in greenhouses. Powdery mildew is a persistent disease problem of lentil plants in greenhouses (Beniwal et al. 1993), and poses a threat to precious breeding materials such as F1 plants. Infections by powdery mildews typically result in small white colonies on leaf surfaces (Fig. 4.1b). Lesions expand to cover entire leaf surfaces and pods. Mycelial growth and conidial production can be especially extensive at flowering. In case of severe 4 Applications of Molecular Markers and DNA Sequences 85 infections, leaves become chlorotic, then curled and necrotic prior to abscission. Yield decline may result and plants sometimes die (Agrawal and Prasad 1997). Even though powdery mildew symptoms are easily recognized, identification of the species that causes the disease could be problematic (Glawe 2008). Knowing the species identity is important in devising management strategies as different species have different host ranges and different life histories. Identification of powdery mildew fungi relies on morphology of reproductive structures. Powdery mildews reproduce sexually by forming sexual structure chasmothecia (teleomorph) and asexually through conidia (anamorph). Traditional belief is that morphology of teleomorphs is more reliable than morphology of anamorphs. Taxonomy of powdery mildews of legumes is traditionally based on a few teleomorphic features, including chasmothecial appendage morphology (Braun 1995; Braun 1987) and host range. Powdery mildew pathogens that produce chasmothecia with multiple asci and dichotomously branched chasmothecial appendages were grouped into the genus Microsphaera, while otherwise similar, mycelioid appendagebearing species were classified within the genus Erysiphe (Braun 1987). However, recent phylogenetic studies of powdery mildew fungi using ribosomal DNA sequences demonstrated that anamorphic features are more indicative of phylogenetic lineages than are teleomorphic features, and that anamorphic characters are of utility in species determination (Braun and Takamatsu 2000; Cunnington et al. 2003; Glawe 2008). Chasmothecial appendages traditionally used to distinguish genera are now used to distinguish species (Braun and Takamatsu 2000). However, teleomorphic state is not always available and most of the time it forms when plants are senescent late in the growing season or does not form at all. It prevents timely detection and identification of the pathogen species. Even though abundant conidia are produced early in the disease development, there are only few anamorphic characters available (such as morphology and dimensions of conidia and conidiophores) to describe species and most of them overlap among closely related species. For example, conidia shape and sizes of E. pisi and E. trifolii are very similar and overlap considerably. Likewise, it is not reliable to use host ranges to identify powdery mildew species because many of them have broad and overlapping host ranges (Amano 1986). Accurate determination of the pathogen species is very important not only for managing the disease, but also in plant breeding programs because different resistance genes may confer resistance to different pathogen species (Epinat et al. 1993). In some instances several powdery mildew species have been reported to occur together on the same host (Epinat et al. 1993; Glawe et al. 2004; Mmbaga et al. 2004). Recent advances in molecular techniques have made it possible to investigate the species level identification of lentil powdery mildew pathogens. Use of molecular characters, especially ITS sequence data, has given promising results for species determination in some powdery mildews (Braun and Takamatsu 2000; Cunnington et al. 2003; Mmbaga et al. 2004; Takamatsu et al. 2002). Powdery mildew of lentil is reported to be caused by two Erysiphe species, E. pisi (Amano 1986), a common pathogen of pea, and E. diffusa (Banniza et al. 2004), a 86 E.N. Njambere et al. Fig. 4.4 Chasmothecium and its appendages of E. trifolii formed on an infected lentil plant. Highly branched chasmothecial appendages (a) and long flexuous nature of the chasmothecial appendages (b) pathogen of soybean. The two species differ in conidia sizes. E. pisi produces conidia larger than those of E. diffusa. The major difference between the two species is that E. diffusa produces chasmothecial appendage with highly branched apices, whereas E. pisi produces mycelioid appendages. Powdery mildew is a frequent and serious disease of lentil plants in our greenhouses, but the species identity is not known. We observed that the conidia sizes larger than those described for E. diffusa. However, it produced chasmothecial appendages with regularly branched apices (Fig. 4.4), raising the possibility that it could be E. diffusa. The contradiction between the anamorphical characters and the teleomorphic characters gave confusion about the species identity. In order to ascertain the species identity of the powdery mildew fungus on lentil plants, we analyzed sequences of rDNA ITS region which led to the discovery of a new species of lentil powdery mildew. 4.3.1 Sample Collection and DNA Sequencing Four samples of powdery mildews were collected from three different greenhouses over a 3-year period and an additional sample from the field was included in this study. Because E. diffusa is also a suspect species, a sample of E. diffusa from wild soybean (Glycine spp., kindly provided by Dr. Randall Nelson of USDA ARS, Urbana, Illinois, USA) was also included for comparison. Total DNA was isolated from conidia and/or mycelia from infected lentil plants using FastDNA1 kit described by Chen et al. (1999). PCR amplification of the ITS region from each sample was performed using the primers ITS1 and ITS4 (White et al. 1990), or 4 Applications of Molecular Markers and DNA Sequences 87 Erysiphe- specific primers that we designed on the basis of conserved sequences of the ITS region of Erysiphe spp., EryF (50 TACAGAGTGCGAGGCTCAGTCG30 ) and EryR (50 GGTCAACCTGTGATCCATGTGACTGG30 ) (Attanayake et al. 2009). Amplified DNA fragments were first cloned into plasmid pCR2.1TOPO (Invitrogen Crop, Calsbad, CA). Plasmids containing inserts were verified by restriction digestion. The inserts were sequenced from both strands using one of the six primers: EryF, EryR, ITS1, ITS4, M13F, and M13R at the Sequencing Core Facility of Washington State University. 4.3.2 Sequence Analysis All the ITS sequences of lentil powdery mildews collected from greenhouses and the field used in this study were identical to one another, but they differed in 18 nucleotide positions from the sequence of E. diffusa from a wild soybean Glycine sp. (Fig. 4.5). Sequences were used in BLASTn searches against the GenBank database (http://www.ncbi.nlm.nih.gov/BLAST) to identify the most similar sequences available in the database. The sequences in the GenBank that showed the highest similarity (one base-pair difference) to the lentil powdery mildew sequence were three identical sequences (AB079853 to AB079855) of E. trifoliilike Oidium sp. from Japan (Okamoto et al. 2002). The sequences in the GenBank that showed the next highest similarity (three base pair differences) were five identical sequences (e.g., AB015913 and AF298542) of E. trifolii (Cunnington et al. 2003; Matsuda et al. 2005; Takamatsu et al. 1999), and another sequence (AB015933) of E. baeumleri (Takamatsu et al. 1999). The ITS sequence of the powdery mildew sample from wild soybean was identical to deposited sequences of E. diffusa in the GenBank. Sequence accessions with high similarity values to the sequences determined in this study were aligned using the ClustalW program and used in phylogenetic analysis using the DNA Parsimony program of the PHYLIP package at http:// bioweb2.pasteur.fr/phylogeny/intro-en.html. Parsimony analysis produced one most parsimonious tree with 113 steps. The sequence of lentil powdery mildew formed a tight cluster (monophyletic group) with sequences of Erysiphe baeumleri, E. trifolii, and E. trifolii- like Oidium spp., and is distantly related to (paraphylectic) E. diffusa. Another powdery mildew sequence from wild soybean specimen, also incorporated in this study, formed a separate clade with E. diffusa sequences in the GenBank. 4.3.3 Species Confirmation As E. trifolii is not previously reported to be a pathogen of lentil, we needed to ascertain that the powdery mildew fungus on lentil is indeed E. trifolii and that it is 88 E.N. Njambere et al. E.trifolii E.diffusa GCCGACCCTCCCACCCGTGTCGATTTGTATCTTGTTGCTTTGGCGGGCCGGGCCGCGTCG 60 GCCGACCCTCCCACCCGTGTCGATTTGTATCTTGTTGCTTTGGCGGGCCGGGCCGCGCTG 60 ********************************************************* * E.trifolii E.diffusa TCGCTGTTCGCAAGGACCTGCGTCGGCCGCCCACC-GGTTTTGAACTGGAGCGCGCCCGC 119 TTGCAGTCCGCATGGACATGCGTCGGCCGCCCCCCCGGTGTTCCACTGGAGCGCGCCCGC 120 * ** ** **** **** ************** ** *** ** **************** E.trifolii E.diffusa CAAAGACCCAACCAAAACTCATGTTGTTTGTGTCGTCTCAGCTTTATTATGAAAATTGAT 179 CAAAGACCCAACCAAAACTCATGTTGTTTGTATCGTCTCAGCTTTATTATGAAAATTGAT 180 ******************************* **************************** E.trifolii E.diffusa AAAACTTTCAACAACGGATCTCTTGGCTCTGGCATCGATGAAGAACGCAGCGAAATGCGA 239 AAAACTTTCAACAACGGATCTCTTGGCTCTGGCATCGATGAAGAACGCAGCGAAATGCGA 240 ************************************************************ E.trifolii E.diffusa TAAGTAATGTGAATTGCAGAATTTAGTGAATCATCGAATCTTTGAACGCACATTGCGCCC 299 TAAGTAATGTGAATTGCAGAATTTAGTGAATCATCGAATCTTTGAACGCACATTGCGCCC 300 ************************************************************ E.trifolii E.diffusa CTTGGTATTCCGAGGGGCATGCCTGTTCGAGCGTCATAACACCCCCTCCAGCTGCCTTTG 359 CTTGGTATTCCGAGGGGCATGCCTGTTCGAGCGTCATAACACCCCCTCCAGCTGCCATTG 360 ******************************************************** *** E.trifolii E.diffusa TGTGGCTGCGGTGTTGGGGCACGTCGCGATGCGGCGGCCCTTAAAGACAGTGGCGGTCCC 419 TGTGGCTGCGGTGTTGGGGCTCGTCGCGATGCGGCGGCCCTTAAAGACAGTGGCGGTTCC 420 ******************** ************************************ ** E.trifolii E.diffusa GGCGTGGGCTCTACGCGTAGTAACTTGCTTCTCGCGACAGAGTGACGACGGTGGCTTGCC 479 GACGTGGGCTCTACGCGTAGTAACTTGCTTCTCGCGACAGAGTGACGACGGTGGCTTGCC 480 * ********************************************************** E.trifolii E.diffusa AGAACACCCCTCTTTTGCTCCAGTCACATGGATCACAGGTTGACC 524 AGAACAACCCTCTTTTGCTCCAGTCACATGGATCACAGGTTGACC 525 ****** ************************************** Fig. 4.5 Alignment of ITS sequences of E. trifolii and E. diffusa determined in this study. An asterisk indicates an identical base pair. There are 18 base-pair differences between the two sequences. The E. trifolii sequence is > 99% similar to previously deposited sequences of E. trifolii in GenBank, and the E. diffusa sequence is identical to previously deposited E. diffusa sequences in GenBank pathogenic on lentil. Three experiments were carried out to confirm that the powdery mildew pathogen of lentil in the US is E. trifolii, and not E. diffusa. First, conidia of E. trifolii were collected from lentil and used in a detached leaf assay to determine the pathogenicity on lentil under controlled conditions. Second, an authentic species of E. trifolii was obtained and compared with the samples from lentil in the US. The experiment showed that E. trifolii does produce long flexuous chasmothecial appendages with regularly branched apices similar to lentil samples (Fig. 4.4a, b). Finally, as E. diffusa is a common pathogen of soybean, soybean genotypes “L84-2237” and “Harosoy” known to be susceptible to E. diffusa were inoculated with conidia of powdery mildew from lentil and grown side by side with infected lentil plants in the greenhouse. The lentil powdery mildew did not infect soybean plants during the entire life cycle of soybean. These evidences strongly support the conclusion that the powdery mildew pathogen found on lentil in US was E. trifolii (Attanayake et al. 2009). 4 Applications of Molecular Markers and DNA Sequences 89 Powdery mildews of plants in the Fabaceae are very complex and have begun to receive more and more attention. Further taxonomic studies are needed because E. trifolii has been regarded as a complex of similar species consisting of E. trifolii, E. baeumleri Magn., and E. asteragali DC. (Braun 1987). The nature of this complex needs to be verified. 4.4 Conclusions Modern molecular techniques have been used for identifying fungi in a wide array of biological science disciplines. In this chapter, we presented two specific examples of how molecular techniques have helped solve practical problems in identifying fungal pathogens of cool season grain legumes. In one case, we used molecular markers (group I introns and ITS sequences) to differentiate S. trifoliorum from a more common and closely related species S. sclerotiorum. This technique of identification allowed us to determine the species identity without the time consuming process of inducing carpogenic germination and ascospore observation. This technique allowed us to identify more than 100 isolates for studies in population genetics of S. trifoliorum. In the second example, using rDNA ITS sequences we were able to identify a new pathogen (E. trifolii) of powdery mildew of lentil. There were some ambiguities in determining the species because the morphology of teleomorph resembled a previously reported species (E. diffusa), but the anamorph is clearly different from E. diffusa. By comparing ITS sequences, examining an authentic specimen of E. trifolii and conducting pathogenicity test of a common host of E. diffusa, we unequivocally determined that the lentil powdery mildew was caused by E. trifolii. In doing so, we actually broadened the taxonomical concept of the species E. trifolii to include regularly branched chasmothecial appendages. Using these two examples, we have shown that modern molecular technology plays an important role and has gained increasing widespread applications in solving practical problems in agriculture. Furthermore, similar to what was found in species of Trichoderma and Hypocrea (Druzhinina et al. 2005), our results showed that the ITS region and the adjacent rDNA could be ideal candidate DNA regions used for developing DNA barcodes for identifying these and related fungal species. References Agrawal SC, Prasad KVV (1997) Diseases of lentil. Science Publishers, Enfield, NH, pp 59–61 Amano K (1986) Host range and geographical distribution of the powdery mildew fungi. Japan Scientific Societies Press, Tokyo, p 543 Attanayake RN, Glawe DA, Dugan FM, Chen W (2009) Erysiphe trifolii causing powdery mildew of lentil (Lens culinaris). Plant Dis 93:797–803 Banniza S, Parmelee JA, Morrall RAA, Tullu A, Beauchamp CJ (2004) First record of powdery mildew on lentil in Canada. Can Plant Dis Surv 84:102–103 90 E.N. Njambere et al. Beniwal SPS, Bayaa B, Weigand S, Makkouk KH, Saxena MC (1993) Field guide to lentil diseases and insect pests. International Center for Agricultural Research in the Dry Areas (ICARDA), Aleppo, Syria Braun U (1987) A monograph of the Erysiphales (Powdery Mildews). Beiheftezur Nova Hedwigia 89:1–700 Braun U (1995) The powdery mildews (Erysiphales) of Europe. Gustav Ficher Verlag, New York, pp 1–307 Braun U, Takamatsu S (2000) Phylogeny of Erysiphe, Microsphaera, Uncinula (Erysipheae) and Cystotheca, Podosphaera, Sphaerotheca (Cystotheceae) inferred from rDNA ITS sequences – some taxonomic consequences. Schlechtendalia 4:1–33 Bretag TW, Mebalds MI (1987) Pathogenicity of fungi isolated from Cicer arietinum (chickpea) grown in northwestern Victoria. Aust J Exp Agric 27:141–148 Chen W, Gray LE, Kurle JE, Grau CR (1999) Specific detection of Phialophora gregata and Plectosporium tabacinum in infected soybean plants. Mol Ecol 8:871–877 Cother EJ (1977) Isolation of important fungi from seeds of Cicer arietinum. Seed Sci Technol 5:593–597 Cunnington JH, Takamatsu S, Lawrie AC, Pascoe IG (2003) Molecular identification of anamorphic powdery mildews (Erysiphales). Australas Plant Pathol 32:421–428 Druzhinina I, Kopchinskiy AG, Komon M, Bissett J, Szakacs G, Kubicek CP (2005) An oligonucleotide barcode for species identification in Trichoderma and Hypocrea. Fungal Genet Biol 42:813–828 Epinat C, Pitrat M, Bertrand F (1993) Genetic analysis of resistance of five melon lines to powdery mildews. Euphytica 65:135–144 Glawe DA (2008) The powdery mildews: a review of the world’s most familiar (yet poorly known) plant pathogens. Annu Rev Phytopathol 46:27–51 Glawe DA, du Toit LJ, Pelter GQ (2004) First report of powdery mildew on potato caused by Leveillula taurica in North America Online. Plant Health Prog. doi:10.1094/PHP-2004-121401-HN Henry T, Iwen PC, Hinrichs SH (2000) Identification of Aspergillus species using internal transcribed spacer regions 1 and 2. J Clin Microb 38(4):1510–1515 Holst-Jensen A, Vaage M, Schumacher T, Johansen S (1999) Structural characteristics and possible horizontal transfer of group I introns between closely related plant pathogenic fungi. Mol Biol Evol 16(1):114–126 Kohn LM (1979) Delimitation of the economically important plant pathogenic Sclerotinia species. Phytopathology 69:881–886 Lieckfeldt E, Seifert KA (2000) An evaluation of the use of ITS sequences in the taxonomy of the Hypocreales. Stud Mycol 45:35–44 Matsuda Y, Sameshima T, Moriura N, Inoue K, Nonomura T, Kakutani K, Nishimura H, Kusakari S, Takamatsu S, Toyoda H (2005) Identification of individual powdery mildew fungi infecting leaves and direct detection of gene expression by single conidium polymerase chain reaction. Phytopathology 95:1137–1143 Mmbaga MT, Klopfenstein NB, Kim MS, Mmbaga NC (2004) PCR-based identification of Erysiphe pulchra and Phyllactinia guttata from Cornus florida using ITS-specific primers. For Pathol 34:321–328 Njambere EN, Chen W, Frate C, Wu BM, Temple S, Muehlbauer FJ (2008) Stem and crown rot of chickpea in California caused by Sclerotinia trifoliorum. Plant Dis 92:917–922 Okamoto J, Limkaisang S, Nojima H, Takamatsu S (2002) Powdery mildew of prairie gentian: characteristics, molecular phylogeny and pathogenicity. J Gen Plant Pathol 68:200–207 Power KS, Steadman JR, Higgins BS, Powers TO (2001) Intraspecific variation within North American Sclerotinia trifoliorum isolates characterized by nuclear small subunit rDNA introns. Proceedings of the XI International Sclerotinia Workshop, Central Science Laboratory, New York, UK 4 Applications of Molecular Markers and DNA Sequences 91 Rehnstrom AL, Free SJ (1993) Methods for the mating of Sclerotinia trifolorium. Exp Mycol 17:236–239 Sanchez-Ballesteros J, Gonzalez V, Salazar O, Acero J, Portal MA, Julian M, Rubio V, Bills GF, Polishook JD, Platas G, Mochales S, Pelaez F (2000) Phylogenetic study of Hypoxylon and related genera based on ribosomal ITS sequences. Mycologia 92(5):964–977 Schneider JHM, Salazar O, Rubio V, Keijer J (1997) Identification of Rhizoctonia solani associated with field grown tulips using ITS rDNA polymorphism and pectic zymograms. Eur J Plant Pathol 103:607–22 Takamatsu S, Hirata T, Sato Y, Nomura Y, Sato Y (1999) Phylogenetic relationships of Microsphaera and Erysiphe section Erysiphe (powdery mildews) inferred from the rDNA ITS sequences. Mycoscience 40:259–268 Takamatsu S, Shin HD, Paksiri U, Limkaisang S, Taguchi Y, Nguyen T-B, Sato Y (2002) Two Erysiphe species associated with recent outbreak of soybean powdery mildew: results of molecular phylogenetic analysis based on nuclear rDNA sequences. Mycoscience 43:333–341 Uhm JY, Fujii J (1983a) Ascospore dimorphism in Sclerotinia trifoliorum and cultural characters of strains from different-sized spores. Phytopathology 73:565–569 Uhm JY, Fujii J (1983b) Heterothallism and mating type mutation in Sclerotinia trifoliorum. Phytopathology 73:569–572 White TJ, Bruns T, Lee S, Taylor J (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ (eds) PCR protocols: a guide to methods and applications. Academic Press, San Diego, pp 315–322 Chapter 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium Kurt Brunner and Robert L. Mach Abstract The determination of fungal biomass in diverse samples plays a key role for questions in the fields of plant pathology and agriculture. Until a decade ago, morphological strain determination and quantification by agar-plating methods were the only techniques to quantify fungal infections. These methods were elaborate and time consuming and the obtained results might not always reflect the biological situation. At the end of the 1990s, numerous groups all over the world started with the molecular characterization of the genus Fusarium and defined several diagnostic sequences in the genome of the most prominent Fusarium species as suitable for the discrimination of isolates. Based on these characteristic sequences originally applied for taxonomic studies, quantitative PCR assays were developed from the turn of the millennium until now. PCR tests for certain species were also developed as well as tests for whole groups producing a particular class of toxins. Until now real-time PCR based Fusarium determinations are applied predominantly in niches in agro-biotechnology. However, to further disseminate the inexpensive and rapid quantitative PCR, the quality of analysis has to be guaranteed by defining several standards concerning the PCR procedure from DNA isolation to data analysis. Additionally, plant breeders and agronomists are familiar with toxin analysis and visual rating systems. So change in people’s mind is necessary to realize the benefits of a novel technique. 5.1 Introduction Fungi of the genus Fusarium are worldwide occurring plant pathogens which cause severe damage to numerous cultivable plants (Weiland and Sundsbak 2000; Mirete et al. 2004; Youssef et al. 2007; Li et al. 2008) with the highest economical K. Brunner and R.L. Mach Institute of Chemical Engineering, Research Area Gene Technology and Applied Biochemistry, Gene Technology Group, Vienna University of Technology, Getreidemarkt 9, A-1060, Vienna e-mail: rmach@mail.zserv.tuwien.ac.at Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_5, # Springer-Verlag Berlin Heidelberg 2010 93 94 K. Brunner and R.L. Mach losses upon infection of maize, wheat, and barley (Windels 2000; Nganje et al. 2004). Fusarium-caused diseases have the potential to destroy crops within several weeks and lead to quality losses in grains in two separate ways: besides the deficit due to reduced yield and kernel size (tombstone kernels), the fungus produces numerous toxic metabolites while colonizing the plant and thereby heavily impairs the quality of cereal grains used for the food and feed industry (Marcia McMullen et al. 1997). The acute or chronic toxicity of Fusarium-released compounds led to the introduction of national limits for mycotoxins in many nations or even to supranational applications of regulations (e.g., limits of the European Community since 2006). Integrated control strategies are indispensable to fight Fusarium diseases in modern agriculture and only the combined effect of (1) planting highly resistant varieties, (2) reasonable crop rotations, and (3) suitable tillage systems can minimize the damages caused by this destructive fungus. Although the knowledge about the Fusarium life-cycle and infection paths increased dramatically, biologic difficulties and economic or ecologic interests prevent achieving sustainable success in controlling these pathogenic fungi: throughout the last decade, plant breeders made substantial progress in the development of Fusarium-tolerant maize and wheat cultivars by identifying genetic regions which are linked to the resistance of plants (Anderson et al. 2001; Medianer 2006; Robertson-Hoyt et al. 2006). However, the resistance against Fusarium is spread over several distinct genetic regions (quantitative trait loci, QTLs) in the plant genome and the breeding of highly resistant plants is elaborate and time-consuming. The severity of infections due to the nonavailability of completely resistant plants is further increased by applying unsuitable cropping systems. The influence of crop rotations on the severity of Fusarium head blight was investigated in several studies (Petcu and Ionibã 1998; Reid et al. 2001) and the economically most lucrative cultivation of alternating maize and wheat specially turned out to be problematic. Fusarium nonhost plants as preceding crops of wheat and maize or as intercrops are often less profitable and thereby of low interest to farmers. Erosion causes a dramatic loss of fertile topsoil, especially in North and South America, and no-till systems have been established in endangered areas to overcome the drawback of conventional farming. Although reduced tillage systems prevent the loss of topsoil, Fusarium inoculum density in the soil increases compared to plow treated fields (Steinkellner and Langer 2004). According to the data of the United Nations Food and Agriculture Organization, more than 250,000 km2 of US farmland are cultivated with no-till methods, closely followed by Brazil, Argentina, and Canada. In addition to the above mentioned factors other uncontrollable factors like the weather conditions at time periods crucial for infections play a key role in affecting the produce. Due to these complex interacting criteria no management systems could be established allowing a complete prevention of Fusarium infection of field crops to avoid mycotoxins in the food and feed chain. As the Fusarium problem in agriculture is not supposed to be overcome within the near future, elaborate monitoring programs have been started in many countries with the aim of observing mycotoxin patterns and fungal population dynamics in selected areas. 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium 95 The traditional method used to isolate and characterize fungi is the cultivation on particular media and microscopic investigations. These conventional identification methods are very time-consuming and have to be performed by skilled personnel to prevent incorrect identification and data interpretation. Moreover, conventional methods used for fungal detection are predominantly semi quantitative by determining the colony forming units from surface sterilized grains plated on particular solid media (Cantalejo et al. 1998; Krysinska-Traczyk et al. 2007). To overcome the drawbacks of culture based identifications, throughout the last decade rapid screening technologies based on DNA identification have been developed and are nowadays well established for Fusarium species. In contrast to conventional detection methods, samples can be tested directly without any elaborate isolation and cultivation steps for a proper classification. These novel identification or discrimination methods include PCR based technologies like DGGE, AFLP, or RFLP and also diagnostic microarrays. All these methods are relatively insensitive to microbial backgrounds and non target organisms. 5.2 Quantification Strategies The broad application of DNA-based identification technologies increased the knowledge on diagnostic DNA fragments of Fusarium species: ITS or IGS sequences (Mishra et al. 2003; Gagkaeva and Yli-Mattila 2004; Konstantinova and Yli-Mattila 2004; Mirete et al. 2004; Yli-Mattila et al. 2004a, b; Jurado et al. 2006; Kulik 2008), mitochondrial DNA (Láday et al. 2004a, b), the b-tubulin encoding gene (Gagkaeva and Yli-Mattila 2004; Mach et al. 2004; Yli-Mattila et al. 2004b), the translation elongation factor gene (Knutsen and Holst-Jensen 2004), and the calmodulin gene (Mulè et al. 2004) were sequenced from numerous Fusarium spp. to allow the design of highly specific PCR primers. On the other hand, the genes from biosynthesis pathways for mycotoxins (Niessen and Vogel 1998; Bakan et al. 2002; Lee et al. 2001; González-Jaén et al. 2004; Nicholson et al. 2004; Bezuidenhout et al. 2006; Baird et al. 2008) were studied accurately to distinguish between producers and nonproducers of certain toxins. The sequence information gained throughout the different molecular taxonomic investigations and population studies forms a broad basis not only for qualitative applications but increasingly for quantitative detection methods. Although different techniques like DGGE, RFLP, and AFLP were applied for the qualitative identification of targets, only the real-time PCR tended to be practicable for quantitative detection methods. 5.2.1 Competitive PCR The first steps toward Fusarium quantification were made around 2000. As realtime PCR technique was still in its infancy, competitive PCR was a common tool to gain information on initial amounts of target DNA in samples. 96 K. Brunner and R.L. Mach Nicholson et al. (1998) designed primer pairs specific to either F. graminearum or F. culmorum from RAPD analysis and optimized the reaction for competitive quantification. The F. culmorum species-specific competitive PCR assay was used to study the effect of inoculum load and timing on stem base disease of winter wheat caused by F. culmorum. The extent of fungal colonization, as measured by fungal DNA content, was greater on plants inoculated earlier in the season and increased with increasing conidial load. The F. graminearum-species specific competitive PCR assay was used to study the colonization of wheat grain by different trichothecene producing and nonproducing isolates of F. graminearum. Edwards et al. (2001) were the first to quantify toxin producing Fusarium species based on the presence of a diagnostic fragment of a key gene for trichothecene biosynthesis. The trichodiene synthase encoding gene tri5 is essential for the first step in trichothecene synthesis and is found only in toxin producing strains. The authors demonstrated a good correlation (r2 = 0.76) between the deoxynivalenol (DON) concentrations in winter wheat and the competitive PCR determined fungal biomass. In contrast to visual ratings which do not frequently correlate which DON concentrations (Hussein et al. 1991), the PCR method turned out to be suitable as a rapid test for toxin concentration. The competitive PCR technique was applied to test the efficiency of seven different fungicides for their potential to reduce Fusarium biomass if compared to untreated controls. 5.2.2 Real-Time PCR With the availability of reliable and affordable real-time PCR cyclers this novel technique found its way into the quantitative detection of plant pathogens including numerous Fusarium species. Real-time PCR has several advantages over competitive PCR: the dynamic range of real-time PCR is usually five to six orders of magnitude rather than two for competitive assays. Furthermore, postreaction processing like gel electrophoresis is unnecessary and thereby the real-time detection of fluorescence saves time and a higher throughput is possible. By using probes with different fluorescent reporter dyes, amplification of more PCR products can be used to detect different strains, polymorphisms, or even single point mutations in a single tube. For the quantification of Fusarium in general, two different approaches for the design of specific qPCR assays were chosen: (1) the determination of one particular Fusarium species with a focus on selectivity to allow the quantification of the target even in a background of highly similar isolates, and (2) the simultaneous quantification of all strains which produce certain toxins like trichothecenes or fumonisins, based on genes involved in the biosynthesis of these metabolites. 5.2.2.1 Species Specific Quantification Reischer et al. (2004) developed a TaqMan based PCR assay for F. graminearum which is the most prevalent species found in moderate climate zones. The method 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium 97 targets the b-tubulin encoding gene tub1 which was isolated from nine Austrian F. graminearum isolates and the sequence was aligned with 144 tub1 sequences from the closely related species F. culmorum, F. poae, F. pseudograminearum, F. sporotrichioides, F. cerealis, and F. lunulosporum to guarantee the specificity of the test. The method developed in this study allows a fast, species-specific identification and quantitation of plant-infections by F. graminearum at very early stages where classical microbiological methods failed to detect the pathogen. The authors demonstrated that five copies of the tub1 gene were sufficient for reliable quantification. The method can be applied on DNA extracted directly from infected plant material and is not affected by any unspecific background of either plant or fungal DNA, even from other pathogens causing head blight. Other TaqMan-based species specific PCR assays were developed for F. graminearum, F. poae, F. culmorum, or F. avenaceum (Waalwijk et al. 2004), the predominant species associated with head blight in Europe. The applied primer pairs were designed from RAPD fragments previously developed (Nicholson et al. 1998) or taken from a previous study (Waalwijk et al. 2003). For all species, the level of quantification was below 1 pg of genomic DNA (what corresponds to 25 fungal genomes) and all assays showed a dynamic range of at least four magnitudes. Based on these quantitative tests, a comprehensive monitoring of the Fusarium community in the Netherlands was performed in 2001 and 2002. Forty wheat fields well distributed all over the country were chosen for analysis and most samples turned out to be infected with a mix of different Fusarium species with F. graminearum occurring as the most prominent one. The authors clearly demonstrated the advantage of reliable PCR systems combined with a high-throughput DNA extraction method over morphologic based monitoring. In contrast to conventional agarplating techniques, the different fungal species can be quantified and usually the analyses are less time-consuming. In general, the PCR results and the data obtained by classic microscopy were quite similar but significant discrepancies were observed for several samples. The authors suppose that these differences were due to the major advantage of DNA based detection methods: the high stability of DNA allows detection of the fungal biomass that was produced during the infection of a kernel, irrespective of whether it stems from live or dead cells. Agar plate-based investigations depend on intact organisms for successful detection. Interestingly, in this study a nonlinear correlation between the fungal DNA and the DON content of samples was observed, which is in contrast to other publications (Schnerr et al. 2002; Fredlund et al. 2008; Yli-Mattila et al. 2008). However, this effect might also result from low extraction efficiencies of the weak infected kernels. Another quantitative PCR assay has been developed recently to quantify Fusarium solani (Li et al. 2008), a soil-borne fungus that infects soybean roots and causes sudden death syndrome. The goal of this study was to develop a real-time quantitative assay to compare the accumulation of genomic DNA among 30 F. solani isolates in inoculated soybean roots. The small subunit of the ribosomal RNA gene was chosen as PCR target and a dual labeled TaqMan probe was designed to ensure the specificity of the method. The authors demonstrated the correlation 98 K. Brunner and R.L. Mach between colony forming units and DNA amount in infected root tissue. This qPCR approach provides useful information for evaluating the aggressiveness of isolates based on the degree of colonization on soybean roots and for selecting F. solani resistant soybean lines. A comprehensive study on real-time PCR detection of different Fusarium species was published only recently (Yli-Mattila et al. 2008). Quantitative tests for F. graminearum and F. poae were developed and the correlation of fungal biomass to the production of certain toxic metabolites was demonstrated. F. poae and F. langsethiae are morphologically almost indistinguishable but the two fungi produce a different spectrum of toxins and thereby microbiologic infection studies can easily lead to misestimating the toxin content of samples. Primers were designed based on a worldwide sequence collection of IGS sequences of F. poae, F. sporotrichioides, F. langsethiae , and F. kyushuense isolates to exclusively amplify the F. poae fragment. Selective primers for F. graminearum were obtained from the sequences of the IGS regions of Finnish F. graminearum isolates. Subsequently, the F. poae specific assay was designed for use in a quantitative multiplex PCR together with a F. langsethiae/F. sporotrichioides specific primer and probe combination. The application of multiplex PCR allows the quantification of all strains in a single run and thus lowers the costs and increases the throughput of analysis. To date this is the only published quantitative multiplex tests for Fusarium species. Additionally, the authors included the tox5 assay (Schnerr et al. 2001) in their monitoring study and compared the respective PCR results with microbiologic determined contamination levels and the toxin content of cereal grains. A correlation was found between the levels of F. poae DNA and nivalenol and enniatins in barley and between the levels of F. graminearum DNA and DON in oats. The correlations between F. poae DNA and nivalenol and F. graminearum DNA and DON levels were significantly higher than those between the mycotoxins and morphologically determined Fusarium contamination levels. 5.2.2.2 Group Specific Quantification Based on the Detection of Toxin Biosynthesis Genes Frequently quantitative PCR tests are based on toxin biosynthesis key-genes as targets for amplification. In contrast to the species specific assays, a whole group of Fusarium spp. that is able to produce a certain class of mycotoxins is quantified in a single run, regardless of their taxonomic belonging. For many applications these types of PCR quantification might be of greater interest, as in general a good correlation between fungal biomass and the toxins in a sample is given. Most assays developed throughout the last decade focus on trichothecene producing species as this class of metabolites turned out to be most relevant for human and animal health due to their acute toxicity. The biosynthetic pathway has previously been studied and one enzyme turned out to be a kind of key step for biosynthesis of all trichothecenes. The trichodiene synthase catalyzes the initial reaction to form trichodiene (Desjardins et al. 1993). The corresponding gene tri5 is 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium 99 located together with ten other genes for this pathway within the trichothecene cluster. This cluster is exclusively present in fungi with the capacity to produce class A and/or class B trichothecenes (Desjardins et al. 1993). The fist tri5 real-time PCR assay was published by Schnerr et al. (2001) with the tox5 primers which were previously used for a group specific qualitative detection of toxin producers (Mulfinger et al. 2000). The test was applied to 30 wheat samples with infections of 0–78% of the kernels. Interestingly, the PCR method showed a high correlation to microbiological quantification of the infection by the plate counting method. However, for several samples the detected Fusarium DNA was much higher or lower than expected from the plate method. This effect is most probably due to the problem that microbiological methods can only count the number of infected kernels but give no hint of the severity of the Fusarium infection on a particular kernel. The intriguing question of the correlation of Fusarium DNA and produced toxins was further investigated in another study by the same authors (Schnerr et al. 2002). Three hundred wheat samples with DON concentrations between not detectable and 34.3 ppm were tested. Data analysis revealed a correlation coefficient of 0.96 between DON content and DNA amounts. In general the correlation appeared to be better at higher infection levels. This might have biological reasons as certain amounts of DON can be metabolized by the plants to DON-3-glucoside (Berthiller et al. 2005) and/or the efficiency of the DNA extraction varies more at lower Fusarium DNA-concentrations. Strausbaugh et al. (2005) investigated the pathogenicity of Fusarium spp. frequently isolated from wheat and barley roots in southern Idaho during four growth-chamber experiments and two field studies. A real-time PCR assay for quantifying the presence of F. culmorum from infected root tissue was developed based on nucleotide sequence for the tri5 gene. In contrast to previous studies (Schnerr et al. 2001, 2002) this test targeted highly variable regions of the tri5 gene to allow specific F. culmorum tests. The TaqMan-based assay is able to quantify F. culmorum in root tissue down to 61 pg of total extracted DNA. Nevertheless, as the tri5 gene is highly conserved among the trichothecene producing strains, the assay also detected F. pseudograminearum and F. graminearum and could not distinguish between these three Fusarium species. Besides the class of trichothecenes, some Fusarium species secrete another threatening class of toxins. Especially in warmer climates the fumonisins produced mainly by F. verticillioides and F. proliferatum are found on maize but rarely on other cereals. The fumonisin biosynthetic genes are clustered (Proctor et al. 2003), and one of these genes, fum1, encodes for a polyketide synthase and was found to be indispensable for fumonisin biosynthesis (B18: Proctor et al. 1999). All quantitative assays for fumonisin producers published up to now focused on the detection of fum1. Bluhm et al. (2004) were the first researchers who used real-time PCR for the group specific detection of fumonisin producers. Nevertheless, instead of proper quantification, the assay was designed and applied to control a certain threshold limit of these strains in barley and maize samples. Another fum1 based quantitative PCR assay was used to study the contamination of more than 420 maize samples from South African farms with fumonisin-toxigenic species (Waalwijk et al. 2008). 100 K. Brunner and R.L. Mach Table 5.1 Overview of quantitative real-time PCR tests for Fusarium species Target species Diagnostic sequence Reference F. graminearum tub1 Reischer et al. (2004) F. avenaceum, F. culmorum, RAPD-derived Waalwijk et al. (2004) F. graminearum, F. poae F. solani small subunit of Li et al. (2008) ribosomal RNA F. graminearum, F. poae IGS Yli-Mattila et al. (2008) tri5 Schnerr et al. (2001) Trichothecene producers F. culmorum, F. graminearum, tri5 Strausbaugh et al. (2005) F. pseudograminearum fum1 Bluhm et al. (2004) Fumonisin producers fum1 Waalwijk et al. (2008) Fumonisin producers The TaqMan based test detected F. verticillioides, F. proliferatum, F. nygami, and F. globosum. Notably, fumonisin nonproducers of F. verticillioides gave no response. The PCR determined DNA amount was compared with the ELISA-based measurements of fumonisin content and a correlation coefficient of 0.87 demonstrated the potential of this method for estimating the toxin content. Table 5.1 shows a summary of previously published quantitative real-time PCR based tests for various plant pathogenic Fusarium species. 5.2.3 DNA-Arrays In many research areas microarray techniques gained a great deal of attention with the growing amount of genetic information of genera and species. Microarrays have the potential to rapidly identify DNA of different origin. In general, the hybridization methods are applied as high throughput systems but the accuracy of quantification is mostly low compared to real-time PCR analysis. A qualitative oligonucleotide array for the differentiation of toxigenic and nontoxigenic Fusarium isolates was developed by Nicolaisen et al. (2005). Until now only a single array based method for the quantitative detection of Fusarium species is available (Kristensen et al. 2007). The capture-oligo sequences for several trichothecene or moniliformin producing groups were designed based on the tef1 sequence and 15 species can be quantified in a single run. Three different capture-probes, each spotted as triplets were included for each species to guarantee the specificity of the assay. The Fusarium chip showed a linear response to diluted Fusarium DNA of more than two magnitudes and the authors proposed a limit of quantification below 16 haploid Fusarium genomes. Barley, oat, wheat, and spelt samples were analyzed morphologically for toxin content and for the species pattern using the array. The results obtained by the novel hybridization method corresponded well with the established analyzes. Although a dilution series was tested to demonstrate the quantification capability of the microarray, all field samples were only tested for the absence or presence of certain species. 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium 5.3 101 Conclusion The quantitative determination of fungal infections is a useful tool in plant pathology for obtaining information on the aggressiveness of different isolates (Li et al. 2008), for Fusarium monitoring in agricultural practice, for testing fungicide efficiencies (Edwards et al. 2001) and has also been used for quality control of wheat, maize, and barley (Schnerr et al. 2001). Microbiologic and morphologic methods are time consuming – usually 7–21 days, depending on the isolates – and furthermore the accuracy of the results depends highly on skilled personnel experienced in microscopic differentiation of strains. Furthermore, the classic plate counting assays to determine colony forming units are often unable to accurately reflect the amount of fungal biomass that earlier led to the infection of plants as they depend on living organisms. In cereal grains the infection and the spread of the pathogen occurred usually 3–7 weeks before harvest. DNA is a relatively stable biomolecule and therefore represents an optimal target for gaining information on past stages of infection. Hence, the quantitative PCR is a brilliant tool for plant pathology related studies because it takes into account living mycelia actively producing toxins and also dead mycelia, which previously led to a certain grade of damage and/or had contributed to the toxin amount found in a sample. Besides microbiologic determination, the analysis of toxins is an established method to gain information about the severity of an infection. Usually expensive HPLC or MS approaches are used to get reliable results. Although these methods are state of the art in mycotoxin analysis, the cost is extremely high and throughput limited. Since more and more countries introduced limits for toxin content in food and feed stuff, a market for commercial ELISA based tests systems was created. However, rapid tests often miss high accuracy and sample preparation is still time consuming. For the control of national mycotoxin limits the precise knowledge of the toxin content is indispensable. In contrast, for plant pathology and agriculture the concentration of a certain metabolite might not be of real interest but until a few years ago toxin tests were the best methods for gaining information about the grade of infection. Or – even worse – taking into account the toxins can only lead to false interpretation of data. Berthiller et al. (2005) demonstrated the high potential of plants to metabolize DON into deoxynivalenol-3-glucoside (D3G) which is reconverted into DON in the gastro-intestinal tract of mammalians. Nowadays there are strong hints that Fusarium resistance breeding heads directly toward the potential of plants to hide fungal toxins like glucosides. Moreover, to save costs these analyzes are only rarely integrated in monitoring projects or in resistance determination of newly bred cultivars. Although several studies demonstrated impressive correlations between Fusarium DNA and the mycotoxin concentration in cereal samples, real-time PCR can only roughly estimate the concentration of these compounds and therefore until now is unsuitable for official controls of food and feed. In contrast to food and feed analysis, the power of this method is found in all niche applications where the fungal biomass plays a key role: (1) quantitative PCR can be used to completely 102 K. Brunner and R.L. Mach substitute morphological analysis from the national Fusarium monitoring projects. In a few PCR runs, numerous species can be detected and for the most prominent strains quantitative assays are available which allow even the analysis of grains with mixed infections. Nevertheless, many studies have been performed with a relatively small subset of reference strains. Therefore over-regional applicability of the tests has to be verified and probably the specificity must be further optimized. (2) Tests for fungicide efficiency can easily be monitored with quantitative PCR. The determination of the DNA amount that caused the disease symptoms in comparison to untreated controls precisely reflects the grade of efficiency of a certain pesticide against Fusarium. The molecular analysis is inexpensive compared to HPLC/MS measurements. (3) Plant breeding and resistance evaluation (including tests for the national approval of cultivars) might be a field with high potential for real-time analysis. Although visual rating systems are well established by plant breeders, this method is elaborate as numerous ears are investigated at several time-points and the grade of infection is plotted against time. A molecular determination of fungal biomass in infected plants is cheaper than any other method and the results are not influenced by any metabolized – and thereby “hidden” – toxins. In general, real-time PCR provides an invaluable potential to facilitate and cheapen analysis in fields like agriculture and plant breeding. However, as the method is still rather in its infancy, there are many problems to be solved in the near future. Until now no validated reference DNA is available on the market. Research groups isolate fungal DNA according to different protocols and use different approaches to determine DNA concentration (e.g., fluorimetric or photometric methods) which do not always give the same results. Furthermore DNA extraction protocols for cereal samples vary from study to study and until now no generally accepted reference method has been introduced. Only recently (Fredlund et al. 2008) the efficiencies of different protocols were compared and almost tenfold differences concerning the concentration of fungal DNA and coextracted PCR inhibitors were revealed. Combining the errors derived from inappropriate DNA-standard concentration measurement with the differences introduced by various DNA extraction protocols makes interlaboratory comparisons of results impossible. Until now no studies with ring-trials or even laborious method evaluations (e.g., repeatability, reproducibility, ruggedness etc.) are available. In general, suggestions for quality assurance for real-time PCR in agro-biotechnology made only recently (Lipp et al. 2005) should be considered as a good approach to increase the quality of analysis and raise the confidence of potential users of this novel technique. A change in people’s mindset – away from established methods – will be necessary to gain high acceptance of the quantitative PCR technique as a useful application for particular applications in plant pathology and agriculture. Acknowledgments The present article was generated in the course of a Fusarium research project financed by the Austrian Federal Ministry of Agriculture, Forestry, Environment, and Water Management, project code FP100053. 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium 103 References Anderson JA, Stack RW, Liu S, Waldron BL, Fjeld AD, Coyne C, Moreno-Sevilla B, Fetch J, Mitchell M, Song QJ, Cregan PB, Frohberg RC (2001) DNA markers for Fusarium head blight resistance QTLs in two wheat populations. Theor Appl Genet 102(8):1164–1168 Baird et al (2008) Identification of select fumonisin forming Fusarium species using PCR applications of the Polyketide Synthase gene and its relationship to fumonisin production in vitro. Int J Mol Sci 9(4):554–570 Bakan B, Giraud-Delville C, Pinson L, Richard-Molard D, Fournier E, Brygoo Y (2002) Identification by PCR of Fusarium culmorum strains producing large and small amounts of deoxynivalenol. Appl Environ Microbiol 68(11):5472–5479 Berthiller F, Dall‘Asta C, Schuhmacher R, Lemmens M, Adam G, Krska R (2005) Masked mycotoxins: determination of a deoxynivalenol glucoside in artificially and naturally contaminated wheat by Liquid Chromatography-Tandem Mass Spectrometry. J Agric Food Chem 53:3421–3425 Bluhm BH, Cousin MA, Woloshuk CP (2004) Multiplex real-time PCR detection of fumonisinproducing and trichothecene-producing groups of Fusarium species. J Food Prot 67(3): 536–543 Bezuidenhout CC, Prinsloo M, Van der Walt AM (2006) Multiplex PCR based detection of potential fumonisin-producing Fusarium in traditional African vegetables. Inc Environ Toxicol 21:360–366 Cantalejo MJ, Carrasco JM, Hernandez E (1998) Incidence and distribution of Fusarium species associated with feeds and seeds from Spain. Rev Iberoam Micol 15:36–39 Desjardins AE, Hohn TM, McCormick SP (1993) Trichothecene biosynthesis in Fusarium species: chemistry, genetics, and significance. Microbiol Rev 57(3):595–604 Edwards SG, Progozliev SR, Hare MC, Jenkinson P (2001) Quantification of trichothecene producing Fusarium species in harvested grain by competitive PCR to determine efficiacies of fungicides against Fusarium head blight of winter wheat. Appl Environ Microbiol 67(4):1575–1580 Fredlund E, Gidlund A, Olsen M, Börjesson T, Spliid NHH, Simonsson M (2008) Method evaluation of Fusarium DNA extraction from mycelia and wheat for down-stream real-time PCR quantification and correlation to mycotoxin levels. J Microbiol Methods 73:33–40 Gagkaeva TY, Yli-Mattila T (2004) Genetic diversity of Fusarium graminearum in Europe and Asia. Eur J Plant Pathol 110:551–562 González-Jaén MT, Mirete S, Patiño B, López-Errasquı́n VC (2004) Genetic markers for the analysis of variability and for production of specific diagnostic sequences in fumonisinproducing strains of Fusarium verticillioides. Eur J Plant Pathol 110:525–532 Hussein HM, Baxter M, Andrew IG, Franich RA (1991) Mycotoxin production by Fusarium species isolated from New Zealand maize fields. Mycopathologia 113:506–511 Jurado M, Vázquez C, Sanchis V, González-Jaén MT (2006) PCR-based strategy to detect contamination with mycotoxigenic Fusarium species in maize. Syst Appl Microbiol 29:681–689 Knutsen AK, Holst-Jensen A (2004) Phylogenetic analyses of the Fusarium poae, Fusarium sporotrichioides and Fusarium langsethiae species complex passed on partial sequences of the translation elongation factor-1 alpha gene. Int J Food Prot 95:287–295 Konstantinova P, Yli-Mattila T (2004) IGS-RFLPanalysis and development of molecular markers for identification of Fusarium poae, Fusarium langsethiae, Fusarium sporotrichioides and Fusarium kyushuense. Int J Food Microbiol 95:321–331 Kristensen R, Gauthier G, Berdal KG, Hamels S, Remacle J, Holst-Jensen A (2007) DNA microarray to detect and identify thrichothecene- and moniliformin-producing Fusarium species. J Appl Microbiol 102:1060–1070 Krysinska-Traczyk E, Perkowski J, Dutkiewicz J (2007) Levels of fungi and mycotoxins in the samples of grain and grain dust collected from five various cereal crops in eastern Poland. Ann Agric Environ Med 14:159–167 104 K. Brunner and R.L. Mach Kulik T (2008) Detection of Fusarium tricinctum from cereal grain using PCR assay. J Appl Genet 49(3):305–311 Láday M, Juhász A, Moretti A, Szécsi A, Logrieco A (2004a) Mitochondrial DNA diversity and lineage determination of European isolates of Fusarium graminearum. Eur J Plant Pathol 110:545–550 Láday M, Mulè G, Moretti A, Hamari Z, Juhász A, Szécsi A, Logrieco A (2004b) Mitochondrial DNA variability in Fusarium proliferatum (Gibberella intermedia). Eur J Plant Pathol 110: 563–571 Lee T, Oh D, Kim H, Lee J, Kim Y, Yun S, Lee Y (2001) Identification of deoxynivalenol and nivalenol-producing chemotypes of Gibberella zeae by using PCR. Appl Environ Microbiol 67(7):2966–2972 Li S, Hartman GL, Domier LL, Boykin D (2008) Quantification of Fusarium solani f. sp. glycines isolates in soybean roots by colony-forming unit assays and real-time quantitative PCR. Theor Appl Genet 117:343–352 Lipp M, Shillito R, Giroux R, Spiegelhalter F, Charlton S, Pinero D, Song P (2005) Polymerase chain reaction technology as analytical tool in agricultural biotechnology. J AOAC Int 88(1):136–155 Mach RL, Kullnig-Gradinger CM, Farnleitner AH, Reischer G, Adler A, Kubicek CP (2004) Specific detection of Fusarium langsethiae and related species by DGGE and ARMS-PCR of a b-tubulin (tub1) gene fragment. Int J Food Prot 95:333–339 Marcia McMullen M, Jones R, Gallenberg D (1997) Scab of wheat and barley: a re-emerging disease of devastating impact. Plant Dis 81(12):1340–1348 Medianer T (2006) Breeding wheat and rye for resistance to Fusarium diseases. Plant Breed 116(3):201–220 Mirete S, Vazquez C, Moulè G, Jurado M, Gonzáles-Jaén MT (2004) Differentiation of Fusarium verticillioides from banana fruits by IGS and EF-1a sequence analysis. Eur J Plant Pathol 110:525–523 Mishra PK, Fox RTV, Culham A (2003) Development of a PCR-based assay for rapid and reliable identication of pathogenic Fusaria. FEMS Microbiol Lett 218:329–332 Mulfinger S, Niessen L,Vogel RF (2000) PCR based quality control of toxigenic Fusarium spp. in brewing malt using ultrasonication for rapid sample preparation. Adv Food Sci 22:38–46 Mulè G, Susca A, Stea G, Moretti A (2004) A species specific PCR assay based on the calmodulin partial gene for identification of Fusarium verticillioides, F. roliferatum and F. subglutinans. Eur J Plant Pathol 110:495–502 Nganje WA, Bangsund DA, Leistritz FL, Wilson WW, Tiapo NM (2004) Regional economic impacts of Fusarium head blight in wheat and barley. Rev Agr Econ 26(3):332–337 Nicholson P, Simpson DR, Weston G, Rezanoor HN, Lees AK, Parry DW, Joyce D (1998) Phys Mol Plant Pathol 53:17–37 Nicholson P, Simpson DR, Wilson AH, Chandler E, Thomsett A (2004) Detection and differentiation of trichothecene and enniatin-producing Fusarium species on small-grain cereals. Eur J Plant Pathol 110:503–514 Nicolaisen M, Justesen AF, Thrane U, Skouboe P, Holmstrom K (2005) An oligonucleotide microarray for the identification and differentiation of trichothecene-producing and nonproducing Fusarium species occurring on cereal grain. J Microbiol Meth 62:57–69 Niessen ML, Vogel RF (1998) Group specific PCR-detection of potential trichothecene-producing Fusarium-species in pure cultures and cereal samples. Syst Appl Microbiol 21(4):618–631 Proctor RH, Desjardins AE, Plattner RD, Hohn TM (1999) A polyketide synthase gene required for biosynthesis of fumonisin mycotoxins in Gibberella fujikuroi mating population A. Fungal Genet Biol 27:100–112 Proctor RH, Brown DW, Plattner RD, Desjardins AE (2003) Co-expression of 15 contiguous genes delineates a fumonisin biosynthetic gene cluster in Gibberella moniliformis. Fungal Genet Biol 38:237–249 5 Quantitative Detection of Fungi by Molecular Methods: A Case Study on Fusarium 105 Reischer GH, Lemmens M, Farnleitner AH, Adler A, Mach RL (2004) Quantification of Fusarium graminearum in infected wheat by species specific real-time PCR applying a TaqMan probe. J Microbiol Meth 59:141–146 Robertson-Hoyt LA, Jines MP, Balint-Kurti PJ, Kleinschmidt CE, White DE, Payne GA, Maragos CM, Molnár TL, Holland JB (2006) QTL mapping for Fusarium Ear rot and Fumonisin contamination resistance in two maize populations. Crop Sci 46:1734–1743 Petcu G, Ioniã S (1998) Influence of crop rotation on weed infestion and Fusarium spp. attack, yield and quality of winter wheat. Rom Agric Res 9–10:83–91 Reid LM, Zhu X, Ma BL (2001) Crop rotation and nitrogen effects on maize susceptibility to Gibberella (Fusarium graminearum) ear rot. Plant Soil 237(1):1–14 Schnerr H, Niessen L, Vogel RF (2001) Real-time detection of the tri5 gene in Fusarium by LightCycler-PCR using SYBR Green I for continous fluorescence monitoring. Int J Food Microbiol 71:53–61 Schnerr H, Vogel RF, Niessen L (2002) Correlation between DNA of trichothecene-producing Fusarium species and deoxynivalenol concentrations in wheat samples. Letters Appl Microbiol 35:121–125 Steinkellner S, Langer I (2004) Impact of tillage on the incidence of Fusarium spp. in soil. Plant Soil 267(1–2):13–22 Strausbaugh CA, Overturf K, Koehn AC (2005) Pathogenicity and real-time PCR detection of Fusarium spp. in wheat and barley roots. Can J Plant Pathol 27:430–438 Waalwijk C, de Kastelein P, Fries PM, Kerenyi Z, van der Lee TAJ, Hesselink T, Köhl J, Kema GHJ (2003) Major changes in Fusarium spp. in wheat in the Netherlands. Eur J Plant Pathol 109:743–754 Waalwijk C, van der Heide R, de Vries I, van der Lee T, Schoen C, Corainville GC, Häusler-Hahn I, Kastelein P, Köhl J, Lonnet P, Demarquet T, Kema GHJ (2004) Quantitative detection of Fusarium in wheat using TaqMan. Eur J Plant Pathol 110:481–494 Waalwijk C, Koch SH, Ncube E, Allwood J, Flett B, de Vries I, Kema GHJ (2008) Quantitative detection of Fusarium spp and its correlation with fumonisin content in maize from South African subsistence farmers. World Mycotox J 1(1):39–47 Weiland JJ, Sundsbak JL (2000) Differentiation and detection of sugar beet fungal pathogens using PCR amplification of actin coding sequences and the ITS region oft he rRNA gene. Plant Dis 84(4):475–482 Windels CE (2000) Economic and social impacts of Fusarium head blight: changing farms and rural communities in the Northern Great Plains. Phytopathology 90(1):17–21 Yli-Mattila T, Mach RL, Alekhina IA, Bulat SA, Koskinen S, Kullnig-Gradinger CM, Kubicek CP, Klemsdal SS (2004a) Phylogenetic relationship of Fusarium langsethiae to Fusarium poae and Fusarium sporotrichioides as inferred by IGS, ITS, b-tubulin sequences and UP-PCR analysis. Int J Food Microbiol 95:267–285 Yli-Mattila T, Paavanen-Huhtala S, Konstantinova P, Gagkaeva TY (2004b) Molecular and morphological diversity of Fusarium species in Finland and north-western Russia. Eur J Plant Pathol 110:573–585 Yli-Mattila T, Paavanen-Huhtala S, Jestoi M, Parikka P, Hietanieme V, Gagkaeva T, Sarlin T, Haikara A, Laaksonen S, Rizzo A (2008) Real-time PCR detection and quantification of Fusarium poae, F. graminearum, F. sporotrichioides and F. langsethiae in cereal grains in Finland and Russia. Arch Phytopathol Plant Prot 41(4):243–260 Youssef SA, Maymon M, Zveibil A, Klein-Gueta D, Sztejnberg A, Shalaby AA, Freeman S (2007) Epidemiological aspects of mango malformation disease caused by Fusarium mangiferae and source of infection in seedlings cultivated in orchards in Egypt. Plant Pathol 56:257–263 Chapter 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize Ivan Visentin, Danila Valentino, Francesca Cardinale, and Giacomo Tamietti Abstract Pink and red ear rot of maize are common diseases in temperate cropping zones. These diseases are caused by toxigenic fungi belonging to the genus Fusarium. Economic losses flow from both reduced yield (shriveled grain) and compromised quality (contamination with mycotoxin). Since the etiology of these diseases is complex and the taxonomy of the genus Fusarium is fluid, there has been a rapid evolution of PCR-based assays for the detection and quantification of toxigenic Fusarium spp. in biological material, and for their assignment to the correct phylogenetic species. Following a brief overview of the symptoms and epidemiology of ear rots in maize, we discuss the toxigenicity of the causal agents and their taxonomy, and finally survey the range of DNA-based tools available for the detection, identification, and quantification of Fusarium spp. pathogenic on maize. 6.1 Introduction Maize, like most cereals, can be infected by a range of pathogens, some of which can significantly damage the economic value of the crop. Fusarium spp. infection of maize has been of particular concern in recent years, because several of these pathogens produce toxic metabolites (mycotoxins) which represent significant contaminants of food and feed (Marasas et al. 1984). Although only a few mycotoxins are considered to represent a realistic threat to human or animal health, some, especially the trichothecenes, zearalenone, and fumonisins, are very stable during seed storage and food/feed processing (Widestrand and Pettersson 2001). Despite extensive toxicological studies, their significance to human health remains unclear, and even less understood is the risk of synergistic interactions when two or more I. Visentin, D. Valentino, F. Cardinale, and G. Tamietti DiVaPRA – Plant Pathology, University of Turin, I–10095 Grugliasco, Turin, Italy e-mail: giacomo.tamietti@unito.it Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_6, # Springer-Verlag Berlin Heidelberg 2010 107 108 I. Visentin et al. toxins occur together in food or feed. With the decreased use of fungicides resulting from the move to lower input cropping technologies and organic farming, there is particular need to monitor the presence of these mycotoxins in cereals and cereal products. The (molecular) diagnosis of plant pathogens requires a detailed knowledge of disease etiology and the taxonomy of the causal agents, and these are complex issues for the maize/Fusarium spp. interaction. Therefore, following a short survey of the biochemistry, toxicity, and mechanisms of action of the major Fusarium toxins which can accumulate in maize, this chapter aims to outline the epidemiology of the common Fusarium diseases of temperate maize, and the main taxonomic issues surrounding the Fusarium genus. The focus, however, is on what DNA-based tools are available to quickly and reliably classify Fusarium isolated from maize, and to diagnose and quantify toxigenic strains from field samples. 6.2 Major Fusarium Toxins in Maize 6.2.1 Fumonisins The fumonisins are a group of food-borne carcinogenic mycotoxins produced by F. verticillioides (Saccardo) Nirenberg, F. proliferatum (Matsushima) Nirenberg, and F. nygamai Burgess & Trimboli. They can be fatal to horses, causing extensive necrosis of brain tissue (equine encephalomalacia), to pigs by chronic accumulation of fluid in the lungs (porcine pulmonary oedema), and to rats by necrosis of the liver. They are suspected to be the etiological agent of oesophageal carcinoma in humans (Marasas et al. 1988). Wild type strains of F. verticillioides produce almost exclusively four B-series fumonisins, molecules consisting of two tricarballylic esters attached to a carbon backbone (Bezuidehout et al. 1988; Nelson et al. 1993). Esterification is an essential step in the maturation of the fumonisins, without which the molecule does not display full biological activity (Seefelder et al. 2003). The most common fumonisin present in naturally contaminated maize is B1 (FB1, Fig. 6.1), a molecule with an aminoeicosapentol backbone with two hydroxyl groups esterified with 3-carboxy-1,5-pentanedioic acid. Less oxygenated fumonisins, which occur at levels considerably lower than FB1, are FB2 (lacking the C-10 hydroxyl group), FB3 (lacking the C-5 hydroxyl group), and FB4 (lacking both OH 1 2 NH2 3 R1 4 5 OR3 6 7 8 9 10 R2 Fig. 6.1 Structure of the B-series fumonisins 11 12 CH3 13 14 OR3 15 16 CH3 17 18 19 20 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 109 groups) (Powell and Plattner 1995). The backbone is formed from the reaction between a C18 polyketide chain and one amino acid. Isotope feeding experiments have shown that C3–C20 of the backbone are derived from acetate, and that the amino group and C-1 and C-2 are derived from alanine (Blackwell et al. 1996; Branham and Plattner 1993). The effect of fumonisin on the plant host is not well understood. Some inhibitory effects of FB1 on the growth of maize callus cells have been reported (Van Asch et al. 1992), and both the induction of foliar symptoms on sweet corn hybrids and the inhibition of H+-ATPase activity have been documented (Glenn et al. 2008, Gutierrez-Najera et al. 2005). FB1 can induce cell death in a fashion reminiscent of the hypersensitive response also in Arabidopsis thaliana, as a result of which FB1 has been used by some researchers to study defence-related cell death signaling events in this model plant (Asai et al. 2000). It has been suggested that FB1 acts differently in A. thaliana than in other plant species (Chivasa et al. 2005). However, the wide range of genetic, genomic, and physiological tools developed in A. thaliana has ensured that much of the exploration of the activity of the fumonisins on the plant host has been carried out in this model plant. A natural primary hypothesis is that the toxicity of fumonisin to plant cells is, as it is for animal cells, associated with the inhibition of ceramide synthase. The structural similarity of FB1 to sphingosin has been established as being the basis for its disruption of sphingolipid metabolism, thereby perturbing various membrane functions and leading to cell death (Abbas et al. 1994; Riley et al. 1996; Williams et al. 2006). An alternative, but nonexclusive hypothesis is that FB1 perturbs plasma membrane functionality by interfering with the activity of H+-ATPase, which is the target for several fungal toxins and elicitors (Marra et al. 1996; Wevelsiep et al. 1993). Gutiérrez-Nàjera and coworkers identified FB1 as a potent inhibitor of plasma membrane H+-ATPase in the maize embryo. FB1 has high affinity for this enzyme, and inhibition is uncompetitive (Gutierrez-Najera et al. 2005). Uncompetitive inhibition is the most effective form of inhibition, but is rather rare in nature, possibly because of the risk it poses for metabolism (CornishBowden 1986). A twofold toxicity mechanism has therefore been proposed for FB1 and its homologues: one is indirect, and acts by raising the level of endogenous sphingoid compounds present through their action on sphinganine N-acyltransferase; the second is direct, acting by uncompetitive inhibition of H+-ATPase. 6.2.2 Trichothecenes The trichothecenes are a group of epoxysesquiterpene molecules which can be conveniently classified, on the basis of their chemical structure, into three types: type A, of which T-2 is an example, type B (such as deoxynivalenol – DON), and the macrocyclic trichothecenes, which are not produced by Fusarium spp. (Fig. 6.2 and not shown). The trichothecenes inhibit eukaryotic translation (McLaughlin et al. 1977), and act as virulence factors in the wheat/Fusarium interaction 110 I. Visentin et al. a b O O O OH OH O O O O O O O OH OH O Fig. 6.2 Structure of T-2 toxin (a) and DON (b) (Bai et al. 2002; Proctor et al. 1995). T-2 (Fig. 6.2a), which is one of the most acutely poisonous of the Fusarium toxins, is produced by F. acuminatum Ell. Kellerm, F. equiseti (Corda) Sacc, and F. sporotrichioides Scherb. T-2 toxintreated A. thaliana seedlings are stunted and aberrant in their morphology, and microarray analysis in A. thaliana has suggested that the toxin induces a number of defence-related responses, inactivates brassinosteroid synthesis, and generates reactive oxygen species (Masuda et al. 2007). DON (Fig. 6.2b), which is much more common in wheat, barley, oat, rice and maize than T-2, is produced primarily by F. graminearum Schwabe (telomorph Gibberella zeae [Schw.] Petch) and F. culmorum Sacc., the causal agents, respectively, of red ear rot in maize and head blight in wheat. DON is also known as vomitoxin because of its emetic effect. The level of DON contamination in cereals varies from harvest to harvest, and is directly correlated with the presence of F. graminearum and F. culmorum. DON appears to inhibit translation in A. thaliana cells, but unlike T-2, this activity is not associated with the induction of a defence response (Masuda et al. 2007). 6.3 Epidemiology and Etiology of Maize Pink and Red Ear Rot Fusarium spp. that generate pink or red ear rot are classified as belonging to the Liseola and Discolor section, respectively. The former disease is more frequent in hot dry climates, typical of the temperate production zones; the latter predominates in cooler, moister climates (Bottalico 1998; Logrieco et al. 1995). Although the importance of Fusarium diseases of maize has been understood for many years, high levels of genetic resistance have yet to be introduced into commercial hybrids, even though several dominant genes determining resistance to fumonisin-producing Fusarium spp. have been identified in various inbred lines (Clements et al. 2004). 6.3.1 Pink Ear Rot This disease occurs on isolated kernels, groups of kernels, or damaged kernels, and is recognized by the formation of a white to light pink mold (Miller 1994). Maize is 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 111 typically grown as either a continuous monoculture or in short rotations with one or two other crops. As a result, most fields retain maize debris in or on the soil, or in neighboring fields. Such plant residue is the primary source of inoculum (Smith and White 1988), as Fusarium spp. survive well on maize residue, either as mycelium or other survival structures (Sutton 1982). In particular, F. verticillioides can produce thickened surviving hyphae (Nyval and Kommdahl 1968), or colonize senescent tissues of other crops and weeds not considered as true hosts (Parry 1995). These heterothallic Fusarium species sporadically produce perithecia, but sexual reproduction is unlikely to play a significant role in their epidemiology. F. verticillioides, F. proliferatum, and F. subglutinans produce large quantities of micro- and macroconidia on crop residues, which act as the most important source of inoculum (Smith and White 1988). Microconidia are typically the more numerous and more easily wind-dispersed than macroconidia. Insects can also play a significant role in inoculum dispersion (Dowd 1998; Gilbertson et al. 1986); in Europe, Ostrinia nubilalis (the corn borer) is the most effective of such vectors, its larvae acquiring the spores from leaf surfaces and transporting them into the kernels (Sobek and Munkvold 1999). Silk infection is also important in the development of symptomless colonization and pink ear rot, especially where insect damage is limited (Desjardins et al. 2002; Munkvold and Desjardins 1997; Nelson et al. 1992). The influence of systemic infection from seed-borne inoculum on the development of pink ear rot is disputed. F. verticillioides can systemically and asymptomatically colonize maize from the infected seed or the root, resulting in invasion of the kernels (Foley 1962; Munkvold et al. 1997). Although there may be certain environmental conditions which favor systemic transmission, kernel infection via this route appears to be only of minor importance under standard field conditions (Desjardins et al. 1998; Munkvold and Carlton 1997). A combination of host genetic resistance (Clements et al. 2004), pathogen variability (Carter et al. 2002; Melcion et al. 1997), and drought stress (the latter being a common event during the grain-filling period in temperate production areas) interacts to modulate the severity of the disease infection and the accumulation of mycotoxin. Several lines of evidence indicate that drought stress is associated with elevated levels of F. verticillioides infection and fumonisin accumulation in kernels (Marin et al. 2001). 6.3.2 Red Ear Rot Red ear rot of maize is caused by one or more of F. graminearum, F. culmorum, F. equiseti, F. chlamydosporum Wollenw. & Reinking, F. acuminatum, F. semitectum Berk. & Rav., and less frequently by F. heterosporum Nees, F. sporotrichioides, F. avenaceum (Corda ex Fries) Sacc. (telomorph Gibberella avenacea Cook), and F. poae (Peck) Wollenw. All these fungi can heavily colonize either the stalk, resulting in premature plant death, and/or the bract, silk, and grain, resulting in the cob, starting from its tip, becoming covered in a pink or red mold (Abbas et al. 1988; Bottalico et al. 1989; Munkvold 2003a). On small grain cereals, 112 I. Visentin et al. infection is associated with reduced seed germination, seedling death, and head blight (Garcia Júnior et al. 2007; Matusinsky et al. 2008; Osborne and Stein 2007; Xu et al. 2008). During the colonization process, a range of mycotoxins is produced, including zearalenone, zearalenol, DON, 3-acetyl DON, 15-acetyl DON, and T-2, all of which accumulate in kernels and serve to enhance disease development (Desjardins et al. 1993, Ohsato et al. 2007; Rocha et al. 2005; Wang et al. 2006). The epidemiology of the diseases incited by the fungi responsible for maize red ear rot has been widely investigated in the small grain crops, but much of the derived knowledge can be readily transferred to maize. The causal agents survive in the seed in plant debris colonized when senescent or dead, in alternative hosts (other crops or weeds) and as chlamydospores on plant debris, and in the soil (Munkvold 2003a; Parry 1995). Debris from potato, sugar beet, and soybean crops have been identified as the source of inoculum for wheat head blight (Broders et al. 2007; Burlakoti et al. 2007). Seed infection is very common and efficient for these Fusarium spp., and allows a ready dispersal of strains (Gilbert et al. 2005; Guo et al. 2008; Shah et al. 2005). Perithecia and/or conidia are produced from colonized plant debris. Gibberella zeae perithecia differentiate when the temperature falls in the range 16–29 C and the substrate moisture level lies between 0.45 and 1.30 MPa (Dufault et al. 2006; Munkvold 2003a). At lower temperatures, or when precipitation exceeds 5 mm, ascospores are ejected from dehydrated perithecia, and are wind-dispersed over long distances (Broders et al. 2007; Osborne and Stein 2007; Trail et al. 2005). The aerial concentration of ascospores begins to rise between 3 and 5 pm, when relative humidity is at its lowest, and peaks at 9 pm. Compared with the population of ascospores, only small numbers of macroconidia are produced from sporodochia on colonized residues (Inch et al. 2005). The temperature optimum for this process is about 29 C and their dispersal mechanism is similar to that of ascospores (Bergstrom and Shields 2002; Tschanz et al. 1976). Within the temperature range 4–30 C and at 100% relative humidity, 50% of ascospores germinate within 33 h. Germination also succeeds at relative humidity levels as low as 53%, but the percentage of germination decreases with the relative humidity (Beyer and Verreet 2005). Host plants are most susceptible to infection at, and shortly after anthesis (Osborne and Stein 2007), but wheat can be infected up until the hard dough stage. Late infections are associated with significant accumulations of DON but not with a loss in grain weight (Ponte et al. 2007). The critical time for F. graminearum infection of the small grain cereals depends on the flowering habit. Thus, barley cultivars with a gaping flower become susceptible just after anthesis, but closed flowering types are only attacked 10 days after anthesis, and this has a significant influence on the amount of mycotoxin accumulated in the grain (Yoshida et al. 2007). A detailed description of the infection process of maize grains via the silk has been given by Miller and co-workers (Miller et al. 2007). Briefly, after germination, the hyphae penetrate the silk and grow towards the cob, ultimately infecting the developing kernels directly through the silk attachment point or indirectly through the ovary wall. Cobs can also be colonized directly via the seed pedicel or glumes. Rainy and humid conditions are particularly conducive 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 113 for disease development, and air temperature has a selective effect on the identity of the pathogen – thus F. graminearum predominates in warm areas and F. culmorum in temperate or cooler ones (Munkvold 2003b, Osborne and Stein 2007). In a survey of wheat and maize grain infected by Fusarium spp. in NW Italy, a low incidence of wheat head blight (1.2–1.4%), and negligible red ear rot and grain mycotoxin contamination in maize were observed over the years 2005–2007, which were characterized by hot, dry weather. However, wheat head blight was extremely common in 2008 (96% for zero tillage crops, and 45% for cultivated ones), a year in which heavy rain and cool conditions prevailed during May and June. Data on maize ear rot for 2008 in the same area are not available yet (G. Tamietti et al, unpublished data). The dependence on cultivation practice underlines the role of crop residues as a source of inoculum (Schaafsma et al. 2005) although the effect can be attenuated where prevailing winds promote long-distance inoculum dispersal (Osborne and Stein 2007). G. zeae ascospores were abundant in the planetary boundary layer throughout crop seasons independent of the time of day, but highly dependent on the extent of cloudiness (Maldonado-Ramirez et al. 2005). Instead, no significant effect of soil management on the incidence of Fusaria as stem-base pathogens in winter wheat was noted (Matusinsky et al. 2008). In wheat, grain colonization by F. graminearum occurs when the temperature is in the range 15–30 C and when water activity (aW) is between 0.900 and 0.995, whereas DON production occurs in the narrower aW range of 0.95–0.995 (Ramirez et al. 2006). Thus cereal varieties able to dry quickly at maturity are probably less prone to DON contamination. 6.4 General Species Concepts and Species Borders Within the Genus Fusarium Mayden considered species concepts as being either theoretical or operational, with the latter being the most relevant in the context of diagnosis (Mayden 1997). All the three common operational species concepts – morphological, biological, and phylogenetic (respectively, MSC, BSC and PSC) – aim to recognize evolutionary distinct species. The theoretical Evolutionary Species Concept (ESC) defines a species as being “.. a single lineage of ancestor-descendent populations which maintains its identity from other such lineages and which has its own evolutionary tendencies and historical fate” (Wiley 1978). The ESC is not informative for species identification, as it is not associated with particular recognition criteria. In contrast, MSC, BSC, and PSC do specify such criteria, but none of the methods of species recognition derived from morphological, biological, and phylogenetic species recognition (respectively, MSR, BSR and PSR) are able to recognize the point at which an ancestral species split into distinct derived species, because changes in morphology, mating behavior, or gene sequences require the passage of time. Under ESC, species are recognized by MSR, BSR, or PSR, but several 114 I. Visentin et al. examples are available of fungal species in which the borders defined in this way do not fully coincide, and this complicates the elaboration of an unequivocal evolutionary pedigree for this kingdom (Taylor et al. 2000). Phylogenetic analyzes based on variable DNA sequences are thought to be more effective in recognizing species consistent with ESC. PSR performs well because evolutionary changes in gene sequence can be recognized long before any changes in mating behavior or morphology (Taylor et al. 2000). Allelic variation at the DNA sequence level does, however, present a problem for PSR, as it can lead to the artefactual splitting of two con-specific isolates into two distinct species; this is a drawback especially in the case of a fungal species lacking a teleomorph (so BSR cannot be applied). This problem is best overcome by consideration of multiple gene genealogies (genealogical concordance phylogenetic species recognition) (Taylor et al. 2000). The content of the genus Fusarium, created by Link (1809), has been variously modified. The first round of reclassifications organized the genus into 16 sections including 65 species, 55 varieties, and 22 forms (Toussoun and Nelson 1975; Wollenweber and Reinking 1935). The main discriminating criteria between sections were morphological, in particular the presence and shape of microconidia, the presence and position along the hyphae of chlamydospores, and the shape of macroconidia and basal cells. All species, varieties, and forms within a section were further characterized by the color of the stroma, the presence of sclerotia, and the number and dimensions of the macroconidial septa. This approach also relied on the observation of fungal growth on different and specific culture media. In 1983, a simpler classification method was proposed, in which the genus was divided into 12 sections and 30 species, and current taxonomic treatments are based on this concept (Burgess et al. 1994; Nelson et al. 1983). In this chapter only the Discolor and Liseola sections are considered, as all the major maize pathogens belong to one or other of these two groups. 6.4.1 Section Liseola This section includes the fumonisin-producing maize pathogens, and comprises the four morphological species F. moniliforme, F. proliferatum, F. subglutinans, and F. anthophilum according to Nelson et al. (1983). Later this was extended to six (F. anthophilum, F. fujikuroi, F. proliferatum, F. sacchari, F. succisae, and F. verticillioides) by Nirenberg (Nirenberg 1989). G. fujikuroi (Sawada) Ito in Ito e K. Kimura is the teleomorph of several Fusarium species within Section Liseola, in which Hsieh et al. identified the three Mating Populations (MP) A, B, C (Hsieh et al. 1977). Subsequently, Kuhlman identified a fourth species and introduced the terminology MP-A (G. fujikuroi var. moniliformis), MP-B (G. fujikuroi var. subglutinans), MP-C (G. fujikuroi var. fujikuroi), and MP-D (G. fujikuroi var. intermedia) (Kuhlman 1989). Since this time, further MPs have been uncovered: MP-E through MP-K (Geiser et al. 2005; Klaasen and Nelson 1996; Lepoint et al. 2005; Leslie 1991; Nirenberg and O’Donnell 1998; Phan et al. 2004; Zeller et al. 2003). 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize Table 6.1 Biological species (MP) in the Liseola section of the G. fujikuroi species complex (modified from (Leslie and Summerell 2006) MP MP-A MP-B MP-C MP-D MP-E MP-F MP-G MP-H MP-I MP-J MP-K Anamorph Fusarium verticillioides F. sacchari F. fujikuroi F. proliferatum F. subglutinans F. thapsinum F. nygamai F. circinatum F. kunzum F. gaditjirrii F. xylarioides 115 Teleomorph Gibberella moniliformis G. sacchari G. fujikuroi G. intermedia G. subglutinans G. thapsina G. nygamai G. circinata G. konza G. gaditjirrii G. xylarioides In the meanwhile, the outcomes of sexual crosses have been gradually integrated with morphological observations and DNA sequence data. The G. fujikuroi species complex in particular has been subjected to various DNA-based phylogenetic analyzes (see later section) (O’Donnell et al. 2000). Overall, the biological, morphological, phylogenetic approaches have produced largely congruent results, and produced the current definition of 11 species (Table 6.1). As alluded to above, current classifications have been largely founded upon and refined by DNA sequence information, which does not rely on observations of the morphology or sexual fertility of any given isolate. In the G. fujikuroi complex, a number of DNA-based approaches has been deployed for this purpose, including electrophoretic karyotyping, RAPD fingerprinting, RFLP genotyping, and DNA sequence comparisons (Steenkamp et al. 1999, 2001; Voigt et al. 1995; Waalwijk et al. 1996; Xu et al. 1995). Thus for example, Xu et al.’s electrophoretic karyotyping method was able to distinguish six MPs (A-F). Increasingly, unknown isolates are assigned to a species on the basis of the DNA sequence within the ribosomal DNA, calmodulin, b-tubulin, and EF-1 genes (Appel and Gordon 1996; Mirete et al. 2004; Mulé et al. 2004, Steenkamp et al. 1999, 2001; Waalwijk et al. 1996). The ribosomal gene family is composed of a tandem array of 18S, 5.8S, and 28S genes, separated from one another by the internal transcribed spacer (ITS) and the intergenic spacer (IGS) regions. As the sequence of the coding regions is well conserved, universal primers can relatively easily be designed; in contrast, the ITS and IGS regions are highly variable, and it is this sequence polymorphism which is exploited for the discrimination between taxa. 6.4.2 Section Discolor Section Discolor comprises 21 species, according to Gerlach and Niremberg (1982), but only six according to Nelson et al. (Nelson et al. 1994). The species fall into two major clades, one producing type A and the other type B trichothecenes. Here, we consider the maize pathogens F. graminearum and F. culmorum, which are the main trichothecene B-producing species. While F. culmorum has a 116 I. Visentin et al. rather distinctive morphology (white or yellow mycelium; stout, thick-walled, and curved macroconidia of width 4–7 mm and length 25–50 mm; abundant chlamydospores occurring singly, in chains, or in clumps), the species borders of F. graminearum have moved significantly in recent years. F. graminearum has been conventionally recognized by its straight, moderately robust macroconidia produced in almost colorless sporodochia, by the absence of microconidia in the aerial mycelium, and by the production of macroconidia and chlamydospores in vegetative mycelium. One subdivision of the species has been based on differences in the ability to produce perithecia in culture (Burgess et al. 1975, Francis and Burgess 1977). This criterion correlates with different ecological behavior, since strains belonging to group I (which do not produce perithecia in agar culture and occur in arid areas) cause crown rot of wheat, while group II strains, later renamed F. pseudograminearum, produce perithecia in culture, are more toxigenic (DON and ZEA), and cause spikelet disease in wheat and cob rot in maize. Later, Miller et al. recognized three chemotypes (able to produce nivalenol, DON, and other acetylated derivatives to different levels) within the broader species concept of F. graminearum (Miller et al. 1991). O’Donnell et al. presented a phylogeny of F. graminearum based on six genes, which indicated a division into seven lineages (or phylogenetic species) with distinct geographic origins. One of these lineages is the producer of DON and ZEA in North America and northern Europe (O’Donnell et al. 1998). Ward et al. used sequence variation at eight toxin genes from a single cluster to identify a phylogeny which was not congruent with those suggested by other genes; the conclusion drawn was that the acquisition of the toxin genes preceded speciation, and that a distinct genetic mechanism unrelated to recombination had been responsible for the maintenance of chemotypes across the phylogenetic species within the morphospecies F. graminearum (Ward et al. 2002). 6.5 DNA Sequence-Based Diagnosis of Fusarium spp. Pathogenic On Maize: Species Assignment The advent of PCR has opened the way to developing simple diagnostic assays based on unique DNA sequence. Here, we review some of the assays which have been developed to discriminate between the Fusarium species pathogenic on maize – i.e., the major agents of pink ear rot (F. proliferatum, F. subglutinans and F. verticillioides) and red ear rot (F. graminearum and F. culmorum). 6.5.1 Species-Diagnostic Primers for the Causal Agents of Pink Ear Rot The first published PCR primers diagnostic for F. moniliforme (i.e., F. verticillioides sensu Nirenberg 1976) were designed from the sequence of a heat shock 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 117 protein gene, within which a RAPD polymorphism had been identified (Murillo et al. 1998). The Fus1-2 primer pair detected the presence of F. verticillioides DNA in infected plants and soils. Möller et al. developed primers, specific for both F. verticillioides (53-6F/R) and F. subglutinans (61-2F/R), based on sequences of RAPD fragments, and were able to employ these for the analysis of infected maize kernels (Möller et al. 1999). When the specificity of these primer pairs was tested against a range of other Fusarium spp. and a selection of other fungal species, 53-6F/R specifically amplified from template of G. fujikuroi MP-A, and 61-2F/R from MP-E. Template of F. culmorum, F. graminearum, or F. proliferatum amplified weakly and only at low annealing temperatures, and the amplicon size was not as expected. Subsequently, Patino et al. developed the F. verticillioides-specific PCR primer pair VERT1/2, based on the sequence of the ribosomal IGS (Patiño et al. 2004). When tested against a panel of 54 F. verticillioides strains obtained from a range of geographical origins and hosts, the assay proved unable to discriminate F. verticillioides from F. proliferatum isolates from Northern Italy (Visentin et al. 2009). A further set of primer pairs were designed by Mulé et al. from the sequence of the taxonomically informative calmodulin gene; these were specific to F. verticillioides (VER1/2), F. subglutinans (SUB1/2), and F. proliferatum (PRO1/2), and their discriminating ability was confirmed in an analysis of 150 maize isolates, mostly from Europe and USA (Mulé et al. 2004). A F. proliferatumspecific primer pair (Fp3F/4R) has also been designed, based on the IGS sequence (Jurado et al. 2006). This assay was successfully validated by testing a range of Fusarium species, commonly associated with cereals, as well as on other fungal genera and plant material. As for the VERT1/2 assay, F. proliferatum could not be always discriminated from isolates of F. verticillioides from Northern Italy (Visentin et al. 2009). Very recently, the previously designed forward primer VERT1 (Patiño et al. 2004) was used along with a newly developed reverse primer VERT-R based on the intergenic spacer region (IGS) to detect F. verticillioides (Sreenivasa et al. 2008). Finally, a new pair of primers designed on the ITS region and to be used in combination with the ITS1 and ITS4 fungal universal primers was described. Although the ITS region offers no full resolution of the genus Gibberella, it may be very useful for the very practical and tedious task of distinguishing unambiguously these two very similar species (Visentin et al. 2009, White et al. 1990). A full list of the Fusarium-specific primers described here is given in Table 6.2. 6.5.2 Species-Diagnostic Primers for the Causal Agents of Red Ear Rot Several diagnostic PCR assays have been developed for the causal agents of red ear rot, in particular F. graminearum and F. culmorum. The first of these to be published involved the two primer pairs UBC85F/R and OPT18F/R, extracted from informative RAPD profiles (Schilling et al. 1996). Both were tested against 118 I. Visentin et al. Table 6.2 Specific PCR assays for F. proliferatum, F. subglutinans and F. verticillioides Primer pairs Primer sequences Specificity Reference Fus1 50 -cttggtcatgggccagtcaagac-30 F. moniliforme Murillo et al. (1998) Fus2 50 -cacagtcacatagcattgctagcc-30 53-6F 50 -tttacgaggcggcgatgggt-30 F. verticillioides Möller et al. (1999) 53-6R 50 -ggccgtttacctggcttctt-30 61-2F 50 -ggccactcaagaggcgaaag-30 F. subglutinans Möller et al. (1999) 61-2R 50 -gtcagaccagagcaatgggc-30 VERT1 50 -gtcagaatccatgccagaacg-30 F. verticillioides Patiño et al. (2004) VERT2 50 -cacccgcagcaatccatcag-30 VER1 50 -cttcctgcgatgtttctcc-30 F. verticillioides Mulé et al. (2004) VER2 50 -aattggccattggtattatatatcta-30 PRO1 50 -ctttccgccaagtttcttc-30 F. proliferatum Mulé et al. (2004) PRO2 50 -tgtcagtaactcgacgttgttg-30 SUB1 50 -ctgtcgctaacctctttatcca-30 F. subglutinans Mulé et al. (2004) SUB2 50 -cagtatggacgttggtattatatctaa-30 Fp3F 50 -cggccaccagaggatgtg-30 F. proliferatum Jurado et al. (2006) Fp4R 50 -caacacgaatcgcttcctgac-30 VERT1 50 -gtcagaatccatgccagaacg-30 F. verticillioides Patiño et al. (2004) VERT-R 50 -cgactcacggccaggaaacc-30 Sreenivasa et al. (2008) verITS-F 50 -aaatcgcgttccccaaattga-30 F. verticillioides White et al. (1990) ITS4 50 -tcctccgcttattgatatgc-30 Visentin et al. (2009) ITS1 50 -tccgtaggtgaacctgcgg-30 F. proliferatum White et al. (1990) proITS-R 50 -gcttgccgcaagggctcgc-30 Visentin et al. (2009) a substantial collection of Fusarium spp. and other fungal pathogens associated with red ear rot. The UBC85F/R primers selectively amplified DNA of all the F. graminearum strains tested, but in some cases, also weakly amplified a fragment from a template of F. culmorum. The OPT18F/R pair amplified as expected from template of 65 out of 69 isolates of F.culmorum obtained from various countries and continents. Using a similar approach (Nicholson et al., 1998) generated the four primer pairs Fc01F/R (specific for F. culmorum), Fg16F/R and Fg16NF/R (F. graminearum), and Fcg17F/R (both F. culmorum and F. graminearum). Fg16F/R amplified selectively from 19 out of 19 isolates of F. graminearum, generating a polymorphic amplicon. A 400 bp product was amplified from the majority of isolates, but three isolates amplified a 470 bp product, one a 500 bp product, and one a 360 bp product. Fg16NF/R generated a monomorphic 280 bp amplicon from all the F. graminearum isolates, and the assay was completely specific. Fc01F/R generated a 570 bp amplicon from 21 out of 21 isolates of F. culmorum, and did not amplify from a template of F. graminearum. Fcg17F/R amplified a 340 bp fragment from all strains of F. graminearum and F. culmorum. Jurado et al. developed an assay for F. graminearum and F. culmorum based on the IGS sequence, and tested it on a diverse set of Fusarium spp. strains commonly associated with cereals (Jurado et al. 2005). The F. culmorum-specific primer pair (Fcu-F/R) amplified a ~ 200 bp fragment from all F. culmorum samples, while the F. graminearum-specific Fgr-F/ R generated a ~ 500 bp amplicon from all but one of the F. graminearum strains tested. A full list of the red ear rot specific assays described here is given in Table 6.3. 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize Table 6.3 Specific PCR assays for F. graminearum and F. culmorum Primer pairs Primer sequences Specificity UBC85F 50 -gcagggtttgaatccgagac-30 F. graminearum UBC85R 50 -agaatggagctaccaacggc-30 OPT18F 50 -gatgccagaccaagacgaag-30 F. culmorum OPT18R 50 -gatgccagacgcactaagat-30 Fc01F 50 -atggtgaactcgtcgtggc-30 F. culmorum FC01R 50 -cccttcttacgccaatctcg-30 Fg16F 50 -ctccggatatgttgcgtcaa-30 F. graminearum Fg16R 50 -ggtaggtatccgacatggcaa-30 Fcg17F 50 -tcgatataccgtgcgatttcc-30 F. culmorum, F. graminearum Fcg17R 50 -tacagacaccgtcaggggg-30 Fg16NF 50 -acagatgacaagattcaggcaca-30 F. graminearum Fg16NR 50 -ttc ttt gac atc tgt tca acc ca-30 Fcu-F 50 -gactatcattatgcttgcgagag-30 F. culmorum Fgc-R 50 -ctctcatataccctccg-30 Fgr-F 50 -gttgatgggtaaaagtgtg-30 F. graminearum Fgc-R 50 -ctctcatataccctccg-30 6.6 119 References Schilling et al. (1996) Schilling et al. (1996) Nicholson et al. (1998) Nicholson et al. (1998) Nicholson et al. (1998) Nicholson et al. (1998) Jurado et al. (2005) Jurado et al. (2005) DNA Sequence-Based Diagnosis of Fusarium spp. Pathogenic on Maize: Toxigenicity The most problematical Fusarium spp. pathogens are those which are toxigenic. Here, we describe available PCR-based methods for identifying and quantifying isolates producing fumonisin and trichothecene from field samples. 6.6.1 Fumonisin-Producing Fusarium spp Sensitive PCR-based methods have been developed to detect the toxigenic species, and especially to identify nonproducing sub-populations or nontoxigenic strains within the toxigenic species. Initially, DNA markers were used in conjunction with phylogenetic methods to distinguish between groups of toxin producers and nonproducers. For example, Gonzalez-Jaèn et al. developed an IGS-RFLP assay which could identify a polymorphism associated with toxigenic strains of F. verticillioides (González-Jaén et al. 2004). Other authors have highlighted the presence of intraspecific polymorphism for such assays. The application of AFLP and IGS/EF-1a sequence variation led to the definition of two F. verticillioides sub-groups, based on a contrast between efficient producers of fumonisin (collected from maize), and nontoxigenic strains (from Central and South American banana fruits) (Mirete et al. 2004; Moretti et al. 2004). The same nontoxigenic population was also exploited by Patiño et al. to generate an IGS-RFLP assay diagnostic for toxigenicity (Patiño et al. 2006). The non-toxigenic isolates were crossable in vitro with MP-A testers (corresponding to G. moniliformis), but showed only about 50% genetic similarity 120 I. Visentin et al. with F. verticillioides strains isolated from maize, and a different chemotoxic profile and virulence on the two hosts (banana and maize). In this case, it was proposed that host specialization had driven the observed genetic drift, which will probably turn into speciation. Several primer pairs were then developed to discriminate toxigenic from non-toxigenic isolates. One of these – diagnostic for F. verticillioides fumonisin-producing strains – was based on sequence of the IGS (Patiño et al. 2004). However, since only nonproducing strains from banana were tested, and all these lack the complete FUM cluster (see below), an assay directed at any of the fumonisin biosynthetic pathway genes would have equally allowed this level of discrimination. Because of the importance of the fumonisin-producing pathogens, some emphasis has been given to elucidating the biosynthetic and regulatory pathway of mycotoxin production. All the fumonisin biosynthetic (FUM) genes characterized to date have been located within a 42.5 kb region of the F. verticillioides genome, in the so-called “FUM cluster” (Brown et al. 2007 Proctor et al. 2003). Gene clusters in this context are distinct from gene families, which are also frequently clustered. The former implies physical proximity, co-regulation, and participation in a common metabolic pathway. The latter, in addition, implies related sequence, as individual members are thought to have evolved by localized duplication and that subsequent divergence. The significance of gene clusters in this sense has long been debated. One hypothesis holds that clustering is associated with gene co-regulation, reminiscent of prokaryotic operons and regulons (Zhang et al. 2004). Alternatively, they may represent an extended form of selfish genes, facilitating simultaneous mobilization of a discrete biosynthetic function for horizontal transfer (Walton 2000). The FUM cluster consists of 16 co-regulated genes on chromosome I, now designated FUM1 (previously named FUM5), FUM2 (previously FUM9), FUM3 (previously FUM12), FUM6-8, FUM10-11, FUM13-19 (Butchko et al. 2003; Proctor et al. 2003, 2006; Seo et al. 2001), and FUM21 (Brown et al. 2007). The role of some of these genes has been deduced by deletion analysis and/or heterologous expression. The deletion of either FUM1, FUM6, or FUM8 blocks the accumulation of all fumonisins, indicating that all are required for fumonisin production (Proctor et al. 1999; Seo et al. 2001). Their role in fumonisin biosynthesis has been inferred from homology to genes of known function and from the analysis of deletion mutants. Several PCR assays have thus been designed to target some of the genes directly involved in mycotoxin biosynthesis, and these have been used in a quantitative mode to correlate the level of fumonisin with the abundance of particular biosynthetic genes (i.e., of the toxigenic fungal strains). In particular, two assays targeting FUM1, which encodes a polyketide synthase, have been applied to a range of Fusarium spp. and other fungal genera. In both cases, the expected amplicon was observed only in fumonisin-producing strains (F. verticillioides and F. proliferatum), but an amplicon of the correct size was also generated from template of the normally non-producing species F. subglutinans and F. thapsinum. López-Errasquı́n et al. have also observed a significant correlation between the expression level of FUM1 and FUM19 and the production of fumonisin B1 (Lopez-Errasquin et al. 2007). Very recently, an additional qPCR test was developed, that targets a 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 121 Table 6.4 PCR assays for the discrimination of fumonisin-producing Fusarium strains Primer pairs Primer sequences References VERTF1 50 -gcgggaattcaaaagtggcc-30 Patiño et al. (2004) VERTF2 50 -gagggcgcgaaacggatcgg-30 FUM5F 50 -gtcgagttgttgaccactgcg-30 Bluhm et al. (2002) FUM5R 50 -cgtatcgtcagcatgatgtagc-30 FUM1for 50 -ccatcacagtgggacacagt-30 Bluhm et al. (2004) FUM1rev 50 -cgtatcgtcagcatgatgtagc-30 PQF1-F 50 -gagccgagtcagcaaggatt-30 Lopez-Errasquin et al. (2007) PQF1-R 50 -agggttcgtgagccaagga-30 PQF19-F 50 -atcagcatcggtaacgcttatga-30 Lopez-Errasquin et al. (2007) PQF19-R 50 -catgtaagttgaggaagcccttgt-30 conserved sequence on the FUM1 gene and gives a good correlation between the estimated total genomic DNA from fumonisin-producing Fusarium species, and fumonisin content in maize kernels (Waalwijk et al. 2008a, b). A list of the primer pairs diagnostic for fumonisin-producing Fusarium spp. reviewed here is given in Table 6.4. 6.6.2 Trichothecene-Producing Fusarium spp The trichothecene pathway is well explored, and several trichothecene biosynthetic genes have been characterized (Desjardins et al. 1993). Bluhm et al. targeted the gene TRI6 to produce two PCR assays diagnostic for trichothecene-producing species. These assays were functional only from template of F. culmorum, F. graminearum, or F. sporotrichioides, all of which are known to be good producers of trichothecenes (Bluhm et al. 2002, 2004). Assays directed at a range of other biosynthetic genes have been designed for the same purpose. One of these targets was TRI5, encoding the catalyst of the isomerization and cyclization of farnyl phosphate to trichodiene (Hohn and Beremand 1989), and this assay was able to detect trichothecene-producing Fusarium spp. both from in vitro cultures and from infected cereal samples (Niessen and Vogel 1998; Niessen et al. 2004). A second assay exploited the trichodiene synthase family member TOX5, and this was not only able to detect the presence of F. graminearum and F. culmorum DNA in cereal samples (Niessen and Vogel 1998), but also its quantitative presence could be correlated with the level of DON (Knoll et al. 2002a, b). Chemotype-specific PCR assays were developed by Ward et al. based on the sequence of the TRI3 and TRI12 genes (Ward et al. 2002). Finally, an assay has recently been developed based on TRI13, and used to explore the toxigenic potential of several Iranian isolates of F. graminearum (Haratian et al. 2008). Several research groups are currently working to establish quantitative PCR methods as a means of correlating the abundance of a toxigenic pathogen in a cereal sample with the amount of trichothecenes present. The necessary primer pairs target either one of the trichothecene biosynthetic genes 122 I. Visentin et al. Table 6.5 PCR assays for the discrimination of trichothecene-producing Fusarium strains Primer pairs Primer sequences References Tri6F 50 -ctctttgatcgtgttgcgtc-30 Bluhm et al. (2002) Tri6R 50 -cttgtgtatccgcctatagtgatc-30 Tri6for 50 -tgatttacatggaggccgaatctca-30 Bluhm et al. (2004) Tri6rev 50 -ttcgaatgttggtgattcatagtcgtt-30 Tox5-1 50 -gctgctcatcactttgctcag-30 Niessen and Vogel (1998) Tox5-2 50 -ctgatctggtcacgctcatc-30 Tri13F 50 -catcatgagacttgtkcrgtttggg-30 Haratian et al. (2008) Tri13R 50 -ttgaaagctccaatgtcgtg-30 TMTrif 50 -cagcagmtrctcaaggtagaccc-30 Halstensen et al. (2006) TMTrir 50 -aactgtayacraccatgccaac-30 Tr5F 50 -agcgactacaggcttccctc-30 Doohan et al. (1999) Tr5R 50 -aaaccatccagttctccatctg-30 FGtubf 50 -ggtctcgacagcaatggtgtt-30 Reischer et al. (2004) FGtubr 50 -gcttgtgtttttcgtggcagt-30 (Halstensen et al. 2006; Schnerr et al. 2002) or the gene encoding beta-tubulin (Reischer et al. 2004). The PCR assays specific to trichothecene-producing Fusarium spp. reviewed here are listed in Table 6.5. 6.7 Conclusion and Future Lines of Research Species definition can be difficult in the fungi, because morphological variation between sibling species is often lacking or difficult to recognize, and because many species lack a known teleomorph. Phylogenetic analyses are therefore particularly valuable to assign isolates to their correct species. We have discussed here a variety of DNA-based tools which allow for a rapid and reliable diagnosis of Fusarium spp. within the Liseola and Discolor sections, and for the detection and quantification of toxin-producing isolates. PCR-based methods are relatively straightforward and quick, but are not totally error-free. Their precision should increase as the informative loci from more isolates of different provenance are sequenced. References Abbas HK, Mirocha CJ, Meronuck RA, Pokorny JD, Gould SL, Kommedahl T (1988) Mycotoxins and Fusarium spp. associated with infected ears of corn in Minnesota. Appl Environ Microbiol 54:1930–1933 Abbas HK, Tanaka T, Duke SO, Porter JK, Wray EM, Hodges L, Sessions AE, Wang E, Merrill AH Jr, Riley RT (1994) Fumonisin- and AAL-toxin-induced disruption of sphingolipid metabolism with accumulation of free sphingoid bases. Plant Physiol 106:1085–1093 Appel DJ, Gordon TR (1996) Relationships among pathogenic and nonpathogenic isolates of Fusarium oxysporum based on the partial sequence of the intergenic spacer region of the ribosomal DNA. Mol Plant Microbe Interact 9:125–138 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 123 Asai T, Stone JM, Heard JE, Kovtun Y, Yorgey P, Sheen J, Ausubel FM (2000) Fumonisin B1induced cell death in arabidopsis protoplasts requires jasmonate-, ethylene-, and salicylatedependent signaling pathways. Plant Cell 12:1823–1836 Bai GH, Desjardins AE, Plattner RD (2002) Deoxynivalenol-nonproducing Fusarium graminearum causes initial infection, but does not cause disease spread in wheat spikes. Mycopathologia 153:91–98 Bergstrom GC, Shields EJ (2002) Atmospheric spore dispersal and regional epidemiology of the Fusarium head blight fungus. Phytopathology 92:S93 Beyer M, Verreet JA (2005) Germination of Gibberella zeae ascospores as affected by age of spores after discharge and environmental factors. Eur J Plant Pathol 111:381–389 Bezuidehout CS, Gerderblom WCA, Gorst-allman CP, Horak RM, Marasas WFO, Spiteller G, Vleggaar R (1988) Structure elucidation of the fumonisins, mycotoxins from Fusarium moniliforme. J Chem Soc Lond Chem Commun 52:743–745 Blackwell BA, Edwards OE, Fruchier A, ApSimon JW, Miller JD (1996) NMR structural studies of fumonisin B1 and related compounds from Fusarium moniliforme. Adv Exp Med Biol 392:75–91 Bluhm BH, Flaherty JE, Cousin MA, Woloshuk CP (2002) Multiplex polymerase chain reaction assay for the differential detection of trichothecene- and fumonisin-producing species of Fusarium in cornmeal. J Food Prot 65:1955–1961 Bluhm BH, Cousin MA, Woloshuk CP (2004) Multiplex real-time PCR detection of fumonisinproducing and trichothecene-producing groups of Fusarium species. J Food Prot 67:536–543 Bottalico A, Logrieco A, Visconti A (1989) Fusarium species and their mycotoxins in infected corn in Italy. Mycopathologia 107:85–92 Bottalico A (1998) Fusarium disease of cereals: species complex and related mycotoxin profiles, in Europe. Eur J Plant Pathol 80:85–103 Branham BE, Plattner RD (1993) Alanine is a precursor in the biosynthesis of fumonisin B1 by Fusarium moniliforme. Mycopathologia 124:99–104 Broders KD, Lipps PE, Paul PA, Dorrance AE (2007) Evaluation of Fusarium graminearum associated with corn and soybean seed and seedling disease in Ohio. Plant Dis 91:1155–1160 Brown DW, Butchko RA, Busman M, Proctor RH (2007) The Fusarium verticillioides FUM gene cluster encodes a Zn(II)2Cys6 protein that affects FUM gene expression and fumonisin production. Eukaryot Cell 6:1210–1218 Burgess LW, Wearing AH, Toussoun TA (1975) Surveys of the fusaria associated with crown rot of weath in eastern Australia. Aust J Agric Res 26:791–799 Burgess LW, Summerell BA, Bullock S, Gott KP, Blackhouse D (1994) Laboratory Manual for Fusarium Research, 3rd edn. Department of Crop Sciences, University of Sydney, Sydney Burlakoti RR, Estrada RJ, Rivera VV, Boddeda A, Secor GA, Adhikari TB (2007) Real-time PCR quantification and mycotoxin production of Fusarium graminearum in wheat inoculated with isolates collected from potato, sugar beet, and wheat. Phytopathology 97:835–841 Butchko RA, Plattner RD, Proctor RH (2003) FUM9 is required for C-5 hydroxylation of fumonisins and complements the meitotically defined Fum3 locus in Gibberella moniliformis. Appl Environ Microbiol 69:6935–6937 Carter JP, Rezanoor HN, Holden D, Desjardins AE, Plattner RD, Nicholson P (2002) Variation in pathogenicity associated with the genetic diversity of Fusarium graminearum. Eur J Plant Pathol 108:573–583 Chivasa S, Ndimba BK, Simon WJ, Lindsey K, Slabas AR (2005) Extracellular ATP functions as an endogenous external metabolite regulating plant cell viability. Plant Cell 17:3019–3034 Clements MJ, Maragos CM, Pataky JK, White DG (2004) Sources of resistance to fumonisin accumulation in grain and Fusarium ear and kernel rot of corn. Phytopathology 94:251–260 Cornish-Bowden A (1986) Why is uncompetitive inhibition so rare? a possible explanation, with implications for the design of drugs and pesticides. FEBS Lett 203:3–6 Desjardins AE, Hohn TM, McCormick SP (1993) Trichothecene biosynthesis in Fusarium species: chemistry, genetics, and significance. Microbiol Rev 57:595–604 124 I. Visentin et al. Desjardins AE, Plattner RD, Lu M, Claflin LE (1998) Distribution of fumonisin in maize ears infected with strains of Fusarium moniliforme that differ in fumonisin production. Plant Dis 82:953–958 Desjardins AE, Munkvold GP, Plattner RD, Proctor RH (2002) FUM1 – a gene required for fumonisin biosynthesis but not for maize ear rot and ear infection by Gibberella moniliformis in field tests. Mol Plant Microbe Interact 15:1157–1164 Doohan FM, Weston G, Rezanoor HN, Parry DW, Nicholson P (1999) Development and use of a reverse transcription-PCR Assay to study expression of Tri5 by Fusarium Species in vitro and in planta. Appl Environ Microbiol 65:3850–3854 Dowd PF (1998) Involvement of arthropods in the establishment of mycotoxigenic fungi under field conditions. In: Sinha K, Bhatnagar D (eds) Mycotoxins in agriculture and food safety. Marcel Dekker, New York, pp 307–350 Dufault NS, Wolf ED, de Lipps PE, Madden LV (2006) Role of temperature and moisture in the production and maturation of Gibberella zeae perithecia. Plant Dis 96:637–644 Foley DC (1962) Systemic infection of corn by Fusarium moniliforme. Phytopathology 52:870–872 Francis RG, Burgess LW (1977) Characteristics of two populations of Fusarium roseum ‘Graminearum’ in eastern Australia. Trans Br Mycol Soc 68:421–427 Garcia Júnior D, Vechiato MH, Menten JOM, Lima MIPM (2007) Influence of Fusarium graminearum on the germination of wheat. Arquivos do Instituto Biológico (São Paulo) 74:157–162 Geiser DM, Ivey ML, Hakiza G, Juba JH, Miller SA (2005) Gibberella xylarioides (anamorph: Fusarium xylarioides), a causative agent of coffee wilt disease in Africa, is a previously unrecognized member of the G. fujikuroi species complex. Mycologia 97:191–201 Gerlach W, Nirenberg HI (1982) The genus Fusarium-A pictorial atlas. Mitt Biol Bundesanst Land- Forstwirtsch Berlin-Dahlem 209:1–406 Gilbert J, Woods SM, Turkington TK, Tekauz A (2005) Effect of heat treatment to control Fusarium graminearum in wheat seed. Can J Plant Pathol 27:448–452 Gilbertson RL, Brown WM, Ruppel EG, Capinera JL (1986) Association of corn stalk rot Fusarium spp. and Western corn rootworm beetles in Colorado. Phytopathology 76:1309–1314 Glenn AE, Zitomer NC, Zimeri AM, Williams LD, Riley RT, Proctor RH (2008) Transformationmediated complementation of a FUM gene cluster deletion in Fusarium verticillioides restores both fumonisin production and pathogenicity on maize seedlings. Mol Plant Microbe Interact 21:87–97 González-Jaén MT, Mirete S, Patiño B, López-Errasquı́n E, Vázquez C (2004) Genetic markers for the analysis of variability and for production of specific diagnostic sequences in fumonisinproducing strains of Fusarium verticillioides. Eur J Plant Pathol 110:525–532 Guo XW, Fernando WGD, Seow-Brock HY (2008) Population structure, chemotype diversity, and potential chemotype shifting of Fusarium graminearum in wheat fields of Manitoba. Plant Dis 92:756–762 Gutierrez-Najera N, Munoz-Clares RA, Palacios-Bahena S, Ramirez J, Sanchez-Nieto S, Plasencia J, Gavilanes-Ruiz M (2005) Fumonisin B1, a sphingoid toxin, is a potent inhibitor of the plasma membrane H + -ATPase. Planta 221:589–596 Halstensen AS, Nordby K, Klemsdal SS, Elen O, Clasen P, Eduard W (2006) Toxigenic Fusarium spp. as determinants of Trichothecene mycotoxins in settled grain dust. J Occup Env Hyg 3:651–659 Haratian M, Sharifnabi B, Alizadeh A, Safaie N (2008) PCR analysis of the Tri13 gene to determine the genetic potential of Fusarium graminearum isolates from Iran to produce Nivalenol and Deoxynivalenol. Mycopathologia 166:109–116 Hohn TM, Beremand PD (1989) Isolation and nucleotide sequence of a sesquiterpene cyclase gene from the trichothecene-producing fungus Fusarium sporotrichiodes. Gene 79:131–138 Hsieh WH, Smith SN, Snyder WC (1977) Mating groups in Fusarium moniliforme. Phytopathology 67:1041–1043 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 125 Inch S, Fernando WGD, Gilbert J (2005) Seasonal and daily variation in the airborne concentration of Gibberella zeae (Schw.) Petch spores in Manitoba. Can J Plant Pathol 27:357–363 Jurado M, Vazquez C, Patino B, Gonzalez-Jaen MT (2005) PCR detection assays for the trichothecene-producing species Fusarium graminearum, Fusarium culmorum, Fusarium poae, Fusarium equiseti and Fusarium sporotrichioides. Syst Appl Microbiol 28:562–568 Jurado M, Vázquez C, Marin S, Sanchis V, González-Jaén MT (2006) PCR-based strategy to detect contamination with mycotoxigenic Fusarium species in maize. Syst Appl Microbiol 29:681–689 Klaasen JA, Nelson PE (1996) Identification of a mating population, Gibberella nygamai sp. nov., within the Fusarium nygamai anamorph. Mycologia 88:965–969 Knoll S, Mulfinger S, Niessen L, Vogel RF (2002a) Rapid preparation of Fusarium DNA from cereals for diagnostic PCR using sonification and an extraction kit. Plant Pathol 51:728–734 Knoll S, Vogel RF, Niessen L (2002b) Identification of Fusarium graminearum in cereal samples by DNA detection test stipestrade. Lett Appl Microbiol 34:144–148 Kuhlman EG (1989) Varieties of Gibberella fujikuroi with anamorph in Fusarium section Liseola. Mycologia 74:759–768 Lepoint PC, Munaut FT, Maraite HM (2005) Gibberella xylarioides sensu lato from Coffea canephora: a new mating population in the Gibberella fujikuroi species complex. Appl Environ Microbiol 71:8466–8471 Leslie JF (1991) Mating populations in Gibberella fujikuroi (Fusarium section Liseola). Phytopathology 81:1058–1060 Leslie JF, Summerell BA (2006) The Fusarium laboratory manual. Blackwell, Iowa, USA Link HF (1809) Observationes in ordines plantarum naturals, Dissetatio I. Mag Ges Naturf Freunde Berlin 3:3–42 Logrieco A, Moretti A, Ritieni A, Bottalico A, Corda P (1995) Occurence and toxigenity of Fusarium proliferatum from preharvest maize ear rot, and associated mycotoxins, in Italy. Plant Dis 79:727–731 Lopez-Errasquin E, Vazquez C, Jimenez M, González-Jaén MT (2007) Real-Time RT-PCR assay to quantify the expression of Fum1 and Fum19 genes from the fumonisin-producing Fusarium verticillioides. J Microbiol Methods 68:312–317 Maldonado-Ramirez SL, Schmale DG, Shields EJ, Bergstrom GC (2005) The relative abundance of viable spores of Gibberella zeae in the planetary boundary layer suggests the role of long-distance transport in regional epidemics of Fusarium head blight. Agr Forest Meteorol 132:20–27 Marasas WF, Kriek NP, Fincham JE, van Rensburg SJ (1984) Primary liver cancer and oesophageal basal cell hyperplasia in rats caused by Fusarium moniliforme. Int J Cancer 34:383–387 Marasas WFO, Nelson PE, Toussoun TA (1988) Reclassification of two important moniliforminproducing strains of Fusarium, NRRL 6022 and NRRL 6322. Mycologia 80:407–410 Marin S, Albareda X, Ramos AJ, Sanchis V (2001) Impact of environment and interactions of Fusarium verticillioides and Fusarium proliferatum with Aspergillus parasiticus on fumonisin B1 and aflatoxins on maize grain. J Sci Food Agric 81:1060–1068 Marra M, Fogliano V, Zambardi A, Fullone MR, Nasta D, Aducci P (1996) The H+-ATPase purified from maize root plasma membranes retains fusicoccin in vivo activation. FEBS Lett 382:293–296 Masuda D, Ishida M, Yamaguchi K, Yamaguchi I, Kimura M, Nishiuchi T (2007) Phytotoxic effects of trichothecenes on the growth and morphology of Arabidopsis thaliana. J Exp Bot 58:1617–1626 Matusinsky P, Mikolasova R, Klem K, Spitzer T, Urban T (2008) The role of organic vs. conventional farming practice, soil management and preceding crop on the incidence of stem-base pathogens on wheat. J Plant Dis Protect 115:17–22 Mayden RL (1997) A hierarchy of species concepts: the denouement in the saga of the species problem. In: Claridge MF, Dawah HA, Wilson MR (eds) Species: the units of biodiversity. Chapman & Hall, London, UK, pp 381–424 126 I. Visentin et al. McLaughlin CS, Vaughan MH, Campbell LM, Wei CM, Stafford ME, Hansen BS (1977) Inhibition of protein synthesis by trichothecenes. In: Rodricks JV, Hesseltine CW, Mehlman MA (eds) Mycotoxins in human and animal health. Pathotox Publishers, Park Forest South, pp 262–272 Melcion D, Cahagnier B, Richard-Molard D (1997) Study of the biosynthesis of fumonisins B1, B2 and B3 by different strains of Fusarium moniliforme. Lett Appl Microbiol 24:301–305 Miller JD, Greenhalgh K, Wang YZ, Lu M (1991) Trichothecene chemotypes of three Fusarium species. Mycologia 83:121–130 Miller JD (1994) Epidemiology of Fusarium diseases of cereals. In: Miller J, Trenholm H (eds) Mycotoxins in grain: compounds other than aflatoxin. Eagan Press, St. Paul, MN, USA, pp 19–36 Miller SS, Reid LM, Harris LJ (2007) Colonization of maize silks by Fusarium graminearum, the causative organism of gibberella ear rot. Can J Bot 85:369–376 Mirete S, Vázquez C, Mulé G, Jurado M, González-Jaén MT (2004) Differentiation of Fusarium verticillioides from banana fruits by IGS and EF-1a sequence analyses. Eur J Plant Pathol 110:515–523 Möller EM, Chelkowski J, Geiger HH (1999) Species-specific PCR assays for the fungal pathogens Fusarium moniliforme and Fusarium subglutinans and their application to diagnose maize ear rot disease. J Phytopathol 147:497–508 Moretti A, Mulè G, Susca A, González-Jaén MT, Logrieco A (2004) Toxin profile, fertility and AFLP analysis of Fusarium verticillioides from banana fruits. Eur J Plant Pathol 110:601–609 Mulé G, Susca A, Stea G, Moretti A (2004) A species-specific PCR assay based on the calmodulin partial gene for the identification of Fusarium verticillioides, F. proliferatum and F. subglutinans. Eur J Plant Pathol 110:495–502 Munkvold GP (2003a) Cultural and genetic approaches to managing mycotoxins in maize. Annu Rev Phytopathol 41:99–116 Munkvold GP (2003b) Epidemiology of Fusarium diseases and their mycotoxins in maize ears. Eur J Plant Pathol 109:705–713 Munkvold GP, Carlton WM (1997) Influence of inoculation method on sistemic Fusarium moniliforme infection of maize plants grown from infected seeds. Plant Dis 81:211–216 Munkvold GP, Desjardins AE (1997) Fumonisins in maize: can we reduce their occurrence. Plant Dis 81:556–565 Munkvold GP, McGee DC, Carlton WM (1997) Importance of different pathways for maize kernel infection by Fusarium moniliforme. Phytophatology 87:209–217 Murillo I, Cavallarin L, San Segundo B (1998) The development of a rapid PCR assay for detection of Fusarium moniliforme. Eur J Plant Pathol 104:301–311 Nelson PE, Toussoun TA, Marasas WFO (1983) Fusarium species: an illustrated manual for identification. The Pennsylvania State University Press, University Park Nelson PE, Plattner RD, Shackelford DD, Desjardins AE (1992) Fumonisin B1 production by Fusarium species other than F. moniliforme in section Liseola and by some related species. Appl Environ Microbiol 58:984–989 Nelson PE, Desjardins AE, Plattner RD (1993) Fumonisins, mycotoxins produced by Fusarium species: biology, chemistry, and significance. Annu Rev Phytopathol 31:233–252 Nelson PE, Dignani MC, Anaissie EJ (1994) Taxonomy, biology, and clinical aspects of Fusarium species. Clin Microbiol Rev 7:479–504 Nicholson P, Simpson DR, Weston G, Rezanoor HN, Lees AK, Parry DW, Joyce D (1998) Detection and quantification of Fusarium culmorum and Fusarium graminearum in cereals using PCR assays. Physiol Mol Plant Pathol 53:17–37 Niessen L, Vogel RF (1998) Group specific PCR-detection of potential trichothecene-producing Fusarium-species in pure cultures and cereal samples. Syst Appl Microbiol 21:618–631 Niessen L, Schmidt H, Vogel RF (2004) The use of tri5 gene sequences for PCR detection and taxonomy of trichothecene-producing species in the Fusarium section Sporotrichiella. Int J Food Microbiol 95:305–319 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 127 Nirenberg H (1976) Investigations on the morphological and biological differentiation in the Fusarium section Liseola. Mitteilungen aus der Biologischen Bundesanstalt fur Land- und Forstwirtschaft. Berlin-Dahlem, 117 pp Nirenberg HI (1989) Identification of Fusaria occurring in Europe on cereals and potatoes. In: Chelkowski J (ed) Fusarium: mycotoxins, taxonomy, and pathology. Elsevier, New York, pp 179–193 Nirenberg HI, O’Donnell K (1998) New Fusarium species and combinations within the Gibberella fujikuroi complex. Mycologia 90:434–458 Nyval RF, Kommdahl T (1968) Individual thickened hyphae as survival structures of Fusarium moniliforme in corn. Phytophatology 58:1704–1707 O’Donnell K, Cigelnik E, Nirenberg HI (1998) Molecular systematics and phylogeography of the Gibberella fujikuroi species complex. Mycologia 90:465–493 O’Donnell K, Nirenberg HI, Aoki T, Cigelnik E (2000) A multigene phylogeny of the Gibberella fujikuroi species complex: detection of additional phylogenetically distinct species. Mycoscience 41:61–78 Ohsato S, Ochiai-Fukuda T, Nishiuchi T, Takahashi-Ando N, Koizumi S, Hamamoto H, Kudo T, Yamaguchi I, Kimura M (2007) Transgenic rice plants expressing trichothecene 3-O-acetyltransferase show resistance to the Fusarium phytotoxin deoxynivalenol. Plant Cell Rep 26:531–538 Osborne LE, Stein JM (2007) Epidemiology of Fusarium head blight on small-grain cereals. Int J Food Microbiol 119:103–108 Parry DW (1995) Fusarium ear blight (scab) in small grain cereals – a review. Plant Pathol 44:207–238 Patiño B, Mirete S, González-Jaén MT, Mulé G, Rodriguez MT, Vazquez C (2004) PCR detection assay of fumonisin-producing Fusarium verticillioides strains. J Food Prot 67:1278–1283 Patiño B, Mirete S, Vázquez C, Jiménez M, Rodrı́guez MT, González-Jaén MT (2006) Characterization of Fusarium verticillioides strains by PCR-RFLP analysis of the intergenic spacer region of the rDNA. J Sci Food Agric 86:429–435 Phan HT, Burgess LW, Summerell BA, Bullock S, Liew ECY, Smith-White JL, Clarkson JR (2004) Gibberella gaditjirrii (Fusarium gaditjirrii) sp. nov., a new species from tropical grasses in Australia. Stud Mycol 50:261–272 Ponte EM, del Fernandes JMC, Bergstrom GC (2007) Influence of growth stage on Fusarium head blight and deoxynivalenol production in wheat. J Phytopathol 155:577–581 Powell RG, Plattner RD (1995) Fumonisins. In: Pelletier SW (ed) Alkaloids, chemical and biological perspectives. Pergamon Press, Oxford, pp 247–278 Proctor RH, Hohn TM, McCormick SP (1995) Reduced virulence of Gibberella zeae caused by disruption of a trichothecene toxin biosynthetic gene. Mol Biol Evol 8:593–601 Proctor RH, Desjardins AE, Plattner RD, Hohn TM (1999) A polyketide synthase gene required for biosynthesis of fumonisin mycotoxins in Gibberella fujikuroi mating population A. Fungal Genet Biol 27:100–112 Proctor RH, Brown DW, Plattner RD, Desjardins AE (2003) Co-expression of 15 contiguous genes delineates a fumonisin biosynthetic gene cluster in Gibberella moniliformis. Fungal Genet Biol 38:237–249 Proctor RH, Plattner RD, Desjardins AE, Busman M, Butchko RA (2006) Fumonisin production in the maize pathogen Fusarium verticillioides: genetic basis of naturally occurring chemical variation. J Agric Food Chem 54:2424–2430 Ramirez ML, Chulze S, Magan N (2006) Temperature and water activity effects on growth and temporal deoxynivalenol production by two Argentinean strains of Fusarium graminearum on irradiated wheat grain. Int J Food Microbiol 106:291–296 Reischer GH, Lemmens M, Farnleitner A, Adler A, Mach RL (2004) Quantification of Fusarium graminearum in infected wheat by species specific real-time PCR applying a TaqMan Probe. J Microbiol Methods 59:141–146 128 I. Visentin et al. Riley RT, Wang E, Schroeder JJ, Smith ER, Plattner RD, Abbas H, Yoo HS, Merrill AH Jr (1996) Evidence for disruption of sphingolipid metabolism as a contributing factor in the toxicity and carcinogenicity of fumonisins. Nat Toxins 4:3–15 Rocha O, Ansari K, Doohan FM (2005) Effects of trichothecene mycotoxins on eukaryotic cells: a review. Food Addit Contam 22:369–378 Schaafsma AW, Tamburic-Ilincic L, Hooker DC (2005) Effect of previous crop, tillage, field size, adjacent crop, and sampling direction on airborne propagules of Gibberella zeae/Fusarium graminearum, fusarium head blight severity, and deoxynivalenol accumulation in winter wheat. Can J Plant Pathol 27:217–224 Schilling AG, Moller ME, Geiger HH (1996) Polymerase chain reaction-based assays for speciesspecific detection of Fusarium culmorum, F. graminearum and F. avenaceum. Phytopathology 86:515–522 Schnerr H, Vogel RF, Niessen L (2002) Correlation between DNA of trichothecene-producing Fusarium species and Deoxynivalenol concentrations in wheat samples. Lett Appl Microbiol 35:121–125 Seefelder W, Humpf HU, Schwerdt G, Freudinger R, Gekle M (2003) Induction of apoptosis in cultured human proximal tubule cells by fumonisins and fumonisin metabolites. Toxicol Appl Pharmacol 192:146–153 Seo JA, Proctor RH, Plattner RD (2001) Characterization of four clustered and coregulated genes associated with fumonisin biosynthesis in Fusarium verticillioides. Fungal Genet Biol 34:155–165 Shah DA, Pucci N, Infantino A (2005) Regional and varietal differences in the risk of wheat seed infection by fungal species associated with fusarium head blight in Italy. Eur J Plant Pathol 115:13–21 Smith DR, White DG (1988) Diseases of corn. In: Sprague GF, Dudley JW (eds) Corn and corn improvement, 3rd edn. American Society of Agronomy, Madison, WI, pp 687–766 Sobek EA, Munkvold GP (1999) European corn borer larvae as vectors of Fusarium moniliforme, causing kernel rot and symptomless infection of maize kernels. J Econ Entomol 92:503–509 Sreenivasa MY, González Jaen MT, Dass RS, Raj APC, Janardhana GR (2008) A PCR-based assay for the detection and differentiation of potential fumonisin-producing Fusarium verticillioides isolated from Indian maize kernels. Food Biotechnol 22:160–170 Steenkamp ET, Wingfield BD, Coutinho TA, Wingfield MJ, Marasas WF (1999) Differentiation of Fusarium subglutinans f. sp. pini by histone gene sequence data. Appl Environ Microbiol 65:3401–3406 Steenkamp ET, Coutinho TA, Desjardins AE, Wingfield BD, Marasas WFO, Wingfield MJ (2001) Gibberella fujikuroi mating population E is associated with maize and teosinte. Mol Plant Pathol 2:215–221 Sutton JC (1982) Epidemiology of wheat head blight and maize ear rot caused by Fusarium graminearum. Can J Plant Pathol 4:195–209 Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS, Fisher MC (2000) Phylogenetic species recognition and species concepts in fungi. Fungal Genet Biol 31:21–32 Toussoun TA, Nelson PE (1975) Variation and speciation in the Fusaria. Annu Rev Phytopathol 13:71–82 Trail F, Gaffoor I, Vogel S (2005) Ejection mechanics and trajectory of the ascospores of Gibberella zeae (anamorph Fusarium graminearum). Fungal Genet Biol 42:528–533 Tschanz AT, Horst RK, Nelson PE (1976) The effect of environment on sexual reproduction of Gibberella zeae. Mycologia 68:327–340 Van Asch MAJ, Rijkenberg FHJ, Coutinho TA (1992) Phytotoxicity of fumonisin B1, moniliformin, and T-2 toxin to corn callus cultures. Phytopathology 82:1330–1332 Visentin I, Tamietti G, Valentino D, Portis E, Karlovsky P, Moretti A, Cardinale F (2009) The ITS region as a taxonomic discriminator between Fusarium verticillioides and Fusarium proliferatum. Mycol Res (in press) 6 DNA-Based Tools for the Detection of Fusarium spp. Pathogenic on Maize 129 Voigt K, Schleier S, Bruckner B (1995) Genetic variability in Gibberella fujikuroi and some related species of the genus Fusarium based on random amplification of polymorphic DNA (RAPD). Curr Genet 27:528–535 Waalwijk C, de Koning JRA, Gams W (1996) Discordant groupings of Fusarium spp. from sections Elegans, Liseola and Dlaminia based on ribosomal ITS1 and ITS2 sequences. Mycologia 88:361–368 Waalwijk C, de Vries IM, Köhl J, Xu X, van der Lee TAJ, Kema GHJ (2008a) Development of quantitative detection methods for Fusarium in cereals and their application. In: Leslie J, Bandyopadhyay R, Visconti A (eds) Mycotoxins: detection methods, management Public Health and Agricultural Trade. CAB International, Wallingford, UK, pp 195–205 Waalwijk C, Koch SH, Ncube E, Allwood J, Flett B, de Vries I, Kema GHJ (2008b) Quantitative detection of Fusarium spp. and its correlation with fumonisin content in maize from South African subsistence farmers. World Mycotoxin J 1:39–47 Walton JD (2000) Horizontal gene transfer and the evolution of secondary metabolite gene clusters in fungi: an hypothesis. Fungal Genet Biol 30:167–171 Wang H, Hwang SF, Eudes F, Chang KF, Howard RJ, Turnbull GD (2006) Trichothecenes and aggressiveness of Fusarium graminearum causing seedling blight and root rot in cereals. Plant Pathol 55:224–230 Ward TJ, Bielawski JP, Kistler HC, Sullivan E, O’Donnell K (2002) Ancestral polymorphism and adaptive evolution in the trichothecene mycotoxin gene cluster of phytopathogenic Fusarium. Proc Natl Acad Sci USA 99:9278–9283 Wevelsiep L, Rupping E, Knogge W (1993) Stimulation of barley plasmalemma H+-ATPase by phytotoxic peptides from the fungal pathogen Rhynchosporium secalis. Plant Physiol 101:297–301 White TJ, Burns T, Lee S, Taylor JW (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfald DH, Sninsky JJ, White TJ (eds) PCR protocol: a guide to methods and application. Academic Press, New York, pp 315–322 Widestrand J, Pettersson H (2001) Effect of time, temperature and solvent on the stability of T-2 toxin, HT-2 toxin, deoxynivalenol and nivalenol calibrants. Food Addit Contam 18:987–992 Wiley EO (1978) The evolutionary species concept reconsidered. Syst Zool 27:17–26 Williams LD, Glenn AE, Bacon CW, Smith MA, Riley RT (2006) Fumonisin production and bioavailability to maize seedlings grown from seeds inoculated with Fusarium verticillioides and grown in natural soils. J Agric Food Chem 54:5694–5700 Wollenweber HW, Reinking OA (1935) Die Fusarien, ihre Beschreibung. Schadwirkung und Bekampfung, Paul Parey, Berlin Xu JR, Yan K, Dickman M, Leslie JF (1995) Electrophoretic karyotypes distinguish the biological species of Gibberella fujikuroi (Fusarium section Liseola). Mol Plant Microbe Interact 8:74–84 Xu XM, Parry DW, Nicholson P, Thomsett MA, Simpson D, Edwards SG, Cooke BM, Doohan FM, Monaghan S, Moretti A, Tocco G, Mule G, Hornok L, Béki E, Tatnell J, Ritieni A (2008) Within-field variability of Fusarium head blight pathogens and their associated mycotoxins. Eur J Plant Pathol 120:21–34 Yoshida M, Kawada N, Nakajima T (2007) Effect of infection timing on Fusarium head blight and mycotoxin accumulation in open- and closed-flowering barley. Phytopathology 97:1054–1062 Zeller K, Summerell BA, Bullock S, Leslie JF (2003) Gibberella konz (Fusarium konzum) sp. nov., a new species within the Gibberella fujikuroi complex from native prairie grasses. Mycologia 95:943–954 Zhang YQ, Wilkinson H, Keller NP, Tsitsigiannis D (2004) Secondary metabolite gene clusters. In: An Z (ed) Handbook of industrial mycology. Marcel Dekker, New York, pp 355–386 Chapter 7 Molecular Detection and Identification of Fusarium oxysporum Ratul Saikia and Narendra Kadoo Abstract Fusarium oxysporum is a ubiquitous inhabitant of soils worldwide and causes diseases such as wilt, yellows, and damping-off in different plant species. Rapid and reliable detection of the pathogen is essential for undertaking appropriate and timely disease management measures. The time-consuming and laborious classical detection methods are now being increasingly replaced by cultureindependent molecular detection techniques, which are much faster, more specific, and sensitive. Molecular techniques like microarrays, whole genome sequencing, DNA barcoding, metagenomics etc. can identify a large number of isolates in a single assay. Some of the emerging tools will also allow complete analysis of developmental processes that are characteristics of the fungus, including the molecular nature of pathogenicity. 7.1 Introduction Fusarium oxysporum Schlechtend. Fr. is an important asexual species complex and is well represented among the soil borne fungi in every type of soil all over the world (Burgess 1981). F. oxysporum includes morphologically indistinguishable pathogenic, nonpathogenic, and even beneficial strains. The pathogenic strains cause diseases such as vascular wilt, yellows, root rot, and damping-off in a wide variety of economically important crops (Beckman and Roberts 1995), while the R. Saikia Biotechnology Division, North-East Institute of Science & Technology, Jorhat, 785006, Assam, India e-mail: rsaikia19@yahoo.com N. Kadoo PMB Group, Biochemical Sciences Division, National Chemical Laboratory, Pune 411008, Maharashtra, India e-mail: ny.kadoo@ncl.res.in Both the authors have contributed equally to the manuscript. Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_7, # Springer-Verlag Berlin Heidelberg 2010 131 132 R. Saikia and N. Kadoo nonpathogenic strains are defined as the strains for which no host plants have been identified (yet) (Lievens et al. 2008). As a species, F. oxysporum probably causes more economic damage to agricultural crops than any other pathogen. In spite of the broad host range of the species as a whole, individual strains usually infect only a single or a few plant species. These individual fungal strains usually show a high level of host specificity and, based on the plant species they can infect, they have been classified into more than 120 formae speciales (Armstrong and Armstrong 1981); for example, F. oxysporum f.sp. ciceri causes wilt only in chickpea. However, some formae speciales such as F. oxysporum f.sp. radicis-cucumerinum, and F. oxysporum f.sp. radicis-lycopersici have broader host ranges, which, apart from infecting cucumber and tomato respectively, can cause root and stem rot on multiple hosts from different plant families (Lievens et al. 2008). Isolates from a particular forma specialis can be further subdivided into physiological races based on cultivar specificity. In addition, based on the ability to form heterokaryons, F. oxysporum strains have been grouped into vegetative compatibility groups (VCGs; Puhalla 1985), and different formae speciales and races may contain multiple VCGs (Katan 1999; Katan and Di Primo 1999). Thus, with regard to effective management of the pathogen, identification below the species level is essential. Identification of F. oxysporum pathotypes is traditionally based on the combination of diagnostic symptoms on the host and the presence of the fungus in the affected tissues (Baayen et al. 2000). However, this classical approach is becoming increasingly challenging because more than one forma specialis may infect a particular host, along with nonpathogenic strains, which are common soil and rhizosphere inhabitants (Edel et al. 2000). Genetic differences among F. oxysporum formae speciales have been evaluated through the analyses of pathogenicity, VCG, chromosomal features, ribosomal DNA (rDNA), mitochondrial DNA (mtDNA), and other molecular markers (Jacobson and Gordon 1990; Puhalla 1985; Katan 1999; Appel and Gordon 1995; O’Donnell et al. 1998; Alves-Santos et al. 1999). However, molecular discrimination of F. oxysporum is complicated by the observation that different isolates classified into a single forma specialis may have independent evolutionary (polyphyletic) origins (O’Donnell et al. 1998; Baayen et al. 2000; Skovgaard et al. 2001; Cramer et al. 2003), and that isolates that belong to different formae speciales may share a common ancestor (monophyletic origin; Kistler 1997). Technological advances in molecular detection methods allow quick and accurate detection and quantification of plant pathogens and these are now being applied to practical problems. The information resulting from such experiments could be used to monitor the level of exposure of the crop to pathogen inoculum and to improve disease control by allowing more rational decisions to be made about the choice and use of fungicides and resistant cultivars. With all these approaches, implementation of appropriate disease management measures requires timely detection and reliable identification of the pathogen and its races. Early and reliable detection is crucial for the containment of the disease and implementation of disease control strategies when they are likely to be most effective. In recent years, the increasing use of molecular methods in fungal diagnostics has emerged 7 Molecular Detection and Identification of Fusarium oxysporum 133 as a possible answer to the problems associated with existing phenotypic identification systems. Here we review the present scenario and emerging advances in molecular identification of plant pathogenic F. oxysporum, and discuss how this knowledge can help in managing the pathogen. 7.2 Earlier Efforts for Identification of Pathogenic Fusarium oxysporum Classically, plant pathogenic fungi were characterized by a series of morphological criteria including cultural characteristics on growth media and diagnostic symptoms on the host along with the presence of the fungus in the affected tissues (Baayen et al. 2000). However, accurate identification of fungi by visual examination of such morphological criteria is very difficult and erroneous. Moreover, these methods have other major limitations such as, reliance on the ability of the fungus to be cultured, time-consuming and laborious nature of identification process, and the requirement for extensive taxonomical knowledge, which complicate timely disease management decisions. Therefore, attempts are being made to replace these methods with molecular identification techniques. As a result, in the last two decades, molecular tools have had a major impact on the identification of plant pathogens. Molecular techniques can avoid many of the drawbacks associated with classical methods of pathogen identification and can also improve our understanding of pathogen detection in different conditions. In general, these techniques are more specific, sensitive, and accurate than traditional methods, and do not demand specialized taxonomical expertise. Today, a wide range of molecular techniques are being applied to accurately identify F. oxysporum isolates (Table 7.1), of which those based on detection of pathogen DNA or RNA are the most predominant. 7.2.1 Identification Using Anonymous Markers Anonymous marker techniques like restriction fragment length polymorphism (RFLP), random amplified polymorphic DNA (RAPD), amplified fragment length polymorphism (AFLP), etc. have been successfully used for identification of F. oxysporum isolates by several workers. 7.2.1.1 Restriction Fragment Length Polymorphism Restriction Fragment Length Polymorphisms (RFLPs) have been extensively used to characterize F. oxysporum isolates and VCGs (Flood et al. 1992; Manicom and Baayen 1993; Fernandez et al. 1994; Mes et al. 1994; Appel and Gordon 1995; 134 R. Saikia and N. Kadoo Table 7.1 Molecular techniques used for identification, detection, or genetic diversity analysis of some formae speciales of Fusarium oxysporum Method of analysis References F. oxysporum forma specialis asparagi Amplified Fragment Length Polymorphism (AFLP) Baayen et al. (2000) albedinis DNA fingerprinting (DF), Restriction Fragment Length Fernandez and Tantaoui (1994); Fernandez et al. (1995) Polymorphism (RFLP), Random Amplified Polymorphic DNA (RAPD), Vegetative Compatibility Grouping (VCG) ciceri RFLP, VCG, RAPD, IGS-RFLP, ISSR Perez-Artes et al. (1995); Honnareddy and Dubey (2006); Singh et al. (2006); Bayraktar et al. (2008) conglutinans DF, isozyme analysis (ISA), Plasmid DNA profile (PDP), Bosland and Williams (1987); Kistler et al. (1987, 1991); RFLP, VCG Kistler and Benny (1989); Hirota et al. (1992) cubense Electrophoretic karyotyping (EK), DF, ISA, RFLP, RAPD, Miao (1990); Ploetz (1990); Kistler et al. (1991); Koenig et al. VCG, AFLP (1993); Boehm et al. (1994); Bentley et al. (1995); O’Donnell et al. (1998); Baayen et al. (2000); Gerlach et al. (2000); Groenewald et al. (2006) cucumerinum DF, RAPD Namiki et al. (1994); Lievens et al. (2007) cyclaminis DF, RFLP, VCG Woudt et al. (1995) dianthi RFLP, VCG, AFLP Manicom et al. (1990); Manicom and Baayen (1993); Baayen et al. (2000) elaeidis RFLP, VCG Flood et al. (1992) gladioli DF, RAPD, VCG, AFLP Mes et al. (1994), Baayen et al. (2000) lini AFLP, VCG Baayen et al. (2000) lycopersici ISA, RAPD, RFLP, VCG, AFLP Elias and Schneider (1991, 1992); Elias et al. (1993); Baayen et al. (2000) melonis DF, RFLP, DNA sequence comparison (DSC), VCG Jacobson and Gordon (1990); Namiki et al. (1994); Appel and Gordon (1995) niveum EK, DF, RFLP, VCG Kim et al. (1993); Namiki et al. (1994) VCG, AFLP RAPD DF, RAPD, VCG radicis-cucumerinum radicis-lycopersici raphani RAPD RFLP, VCG DF, ISA, PLP, RFLP, VCG tulipae vasinfectum VCG, AFLP RAPD, RFLP, VCG, AFLP Baayen et al. (2000) Alves-Santos et al. (2002) Whitehead et al. (1992); Bodker et al. (1993); Grajal-Martin et al. (1993) Lievens et al. (2007) Katan et al. (1991) Bosland and Williams (1987); Kistler et al. (1987,1991); Kistler and Benny (1989); Hirota et al. (1992) Baayen et al. (2000) Fernandez et al. (1994b); Abd-Elsalam et al. (2002a, 2004); Abo et al. (2005) 7 Molecular Detection and Identification of Fusarium oxysporum opuntiarum phaseoli pisi 135 136 R. Saikia and N. Kadoo Baayen et al. 1997; Kistler 1997). Baayen et al. (1998) screened isolates of F. oxysporum from lily (F. oxysporum f.sp. lilii) for pathogenicity, vegetative compatibility, and RFLP patterns, and compared these to reference isolates of the formae speciales gladioli and tulipae. They found that the isolates from Europe and United States shared unique RFLP patterns and belonged to the same VCG. RFLP analysis of Fusarium isolates from carnation by Manicom et al. (1990) and Manicom and Baayen (1993) showed two major VCGs, each characterized by a distinct RFLP pattern. Similarly, Fernandez et al. (1994) used RFLP analysis to identify four ribosomal DNA (rDNA) and seven mitochondrial DNA (mtDNA) haplotypes in F. oxysporum f.sp. vasinfectum, the causal organism of cotton wilt. Attitalla et al. (2004) evaluated isozyme analysis, mtDNA-RFLP, and high performance liquid chromatography (HPLC) to differentiate two morphologically indistinguishable formae speciales of F. oxysporum, lycopersici, and radicis-lycopersici. Although HPLC produced distinct profiles for nonpathogenic and pathogenic isolates, the direct mtDNA-RFLP technique proved to be an efficient diagnostic tool for routine differentiation of lycopersici and radicis-lycopersici isolates (Attitalla et al. 2004). However, although RFLP has been successfully used in many studies to identify Fusarium isolates, due to its labor-intensive nature, elaborate procedure, and the need for high amount of DNA (Garcia-Mas et al. 2000), it is being replaced by polymerase chain reaction (PCR) based techniques. PCR allows rapid detection and identification of pathogens and overcomes most of the limitations of classical approaches. It has revolutionized the detection of pathogens and PCR-based methods are now widely used for identification of a variety of pathogens because of its rapid, sensitive, and specific nature. Many PCRbased approaches have been reported for identification of F. oxysporum isolates and the study of the genetic relationships among them. These fungi have been differentiated using either mycotoxigenic genes, ribosomal DNA, other genes, or unique DNA bands from RAPD analysis (reviewed by Edwards et al. 2002). 7.2.1.2 Random Amplified Polymorphic DNA Random amplified polymorphic DNA (RAPD) is a quick and cost-effective method to detect pathogens and study the genetic similarity or diversity among pathogen populations. The technique has been extensively used to analyze genetic diversity among different F. oxysporum formae speciales, and races (Grajal-Martin et al. 1993; Bentley et al. 1994; Kelly et al. 1994; Manulis et al. 1994; Wright et al. 1996). Paavanen-Huhtala et al. (1999) analyzed 27 F. oxysporum isolates by RAPD and isozyme patterns; however, all the isolates could only be distinguished from each other by RAPD analysis. Mes et al. (1999) screened two races of F. oxysporum f.sp. lycopersici for vegetative compatibility and characterized them using RAPD analysis, and found that the RAPD profiles coincided with the vegetative compatibility groups. The RAPD technique has been used to differentiate a collection of isolates into races corresponding to pathogenicity tests in cotton (Assigbetse et al. 1994) and basil (Chiocchetti et al. 1999; Chiocchetti 2001). Jimenez-Gasco et al. (2001) 7 Molecular Detection and Identification of Fusarium oxysporum 137 identified specific RAPD amplification profiles for F. oxysporum f.sp. ciceri races 0, 1B/C, 5, and 6. Using RAPD-generated DNA probes, Wang et al. (2001) developed a sensitive and specific method for identifying F. oxysporum f.sp. cucumerinum and F. oxysporum f.sp. luffae isolates. After RAPD analysis of 13 formae speciales of F. oxysporum, they selected specific DNA bands as probes and developed forma specialis-specific probes for identification of F. oxysporum f.sp. cucumerinum and F. oxysporum f.sp. luffae isolates by dot blot hybridization. Lievens et al. (2007) developed a robust RAPD marker-based assay to specifically detect and identify the cucumber pathogens F. oxysporum f.sp. cucumerinum and F. oxysporum f.sp. radicis-cucumerinum. Based on the phylogeny of translation elongation factor-1a (TEF-1a), they found that F. oxysporum f.sp. cucumerinum strains were genetically more diverse, while the F. oxysporum f.sp. radiciscucumerinum strains clustered in a separate clade. The developed markers were implemented in an DNA array to enable parallel and sensitive detection and identification of the pathogens in complex samples from diverse origins. However, although the RAPD technique has been successfully used in many studies for detection and identification of F. oxysporum isolates as well as to evaluate the genetic diversity within and among pathogen populations, it suffers from wellknown limitations of poor reproducibility and inter-laboratory transferability. 7.2.1.3 Amplified Fragment Length Polymorphism Amplified fragment length polymorphism (AFLP) has been used in many studies for the analysis of fungal population structure (Majer et al. 1996; Gonzalez et al. 1998; DeScenzo et al. 1999; Purwantara et al. 2000; Zeller et al. 2000). Genetic variation among pathogenic isolates of F. oxysporum was estimated using AFLP markers by several workers (Baayen et al. 2000; Bao et al. 2002; Sivaramakrishan et al. 2002; Groenewald et al. 2006; Stewart et al. 2006). Later, the utility, reproducibility, and efficiency of AFLP technique led to its broader application in the analysis of population diversity and identification of pathogens (Baayen et al. 2000; Abd-Elsalam et al. 2002a, b; Kiprop et al. 2002; Sivaramakrishan et al. 2002; Abdel-Satar et al. 2003; Leslie et al. 2005; Gurjar et al. 2009). The technique was used to examine genetic relationships among isolates of F. oxysporum f.sp. vasinfectum by Abd-Elsalam et al. (2004) and Wang et al. (2006). Gurjar et al. (2009) identified two F. oxysporum f.sp. ciceri races (1 and 2) based on unique AFLP patterns. Sequence characterization of these race-specific AFLP products revealed significant homologies with metabolically important fungal genes. However, as AFLP is relatively costly and has a rather complicated technical procedure, it is being increasingly replaced by simpler PCR-based methods. 7.2.1.4 Simple Sequence Repeats Simple sequence repeats (SSRs), also known as microsatellites, provide a powerful tool for taxonomic and population genetics studies. They have also been used in 138 R. Saikia and N. Kadoo fungal studies because of the high resolution that they provide (Bogale et al. 2005, 2006; Bayraktar et al. 2008). van der Nest et al. (2000) used inter-simple sequence repeat (ISSR) and SSR primers (random amplified microsatellites, RAMS) in PCR to develop SSR markers for F. oxysporum. Barve et al. (2001) assessed the genetic variability in F. oxysporum f.sp. ciceri (Foc) populations prevalent in India using 13 oligonucleotide probes and 11 restriction enzymes. Using the distribution of microsatellite repeats, it was found that races 1 and 4 were closely related as compared to race 2, while race 3 of the pathogen was very distinct. However, as these anonymous marker techniques have several disadvantages, diagnostic DNA fragments identified with these approaches have often been converted into more simple and reliable molecular markers like sequence characterized amplified region (SCAR) or sequence tagged sites (STS). This approach has proven to be effective for the identification of several formae speciales and races of F. oxysporum. For example, Kelly et al. (1998) developed an in planta PCR method to detect isolates of race 5 of Foc in chickpea. The assay using RAPD-derived SCAR markers specifically identified race 5 of the pathogen from infected chickpea plants. Similarly, Jimenez-Gasco and Jimenez-Diaz (2003) sequenced previously identified Foc specific RAPD markers and designed SCAR markers to identify Foc and its four pathogenic races 0, 1A, 5, and 6. The assays were sensitive enough to detect as low as 100 pg of fungal genomic DNA. Based on RAPD analysis, Shimazu et al. (2005) developed three sets of STS markers for specific identification of three races of F. oxysporum f.sp. lactucae. These markers were specific to F. oxysporum f.sp. lactucae and did not amplify DNA from isolates of five other F. oxysporum formae speciales as well as other plant pathogenic fungi, bacteria, or plant materials examined in the study. 7.2.2 Identification Using Sequence-Specific Markers Although the above-mentioned techniques have been successful in accurately identifying the pathogens in many cases, the markers can be localized anywhere in the pathogen genome and often little sequence data are available in public databases for comparison with other sequences. Therefore, extensive screening using a large collection of strains is necessary to validate the robustness of these markers. Lievens et al. (2008) listed specific PCR primers for the detection and identification of several formae speciales and races of F. oxysporum. Such markers that are based on specific DNA sequences in the pathogen genomes could be used for pathogen identification as well as for their phylogenetic studies. 7.2.2.1 ITS and IGS The internal transcribed spacer (ITS) and intergenic spacer (IGS) regions of the ribosomal RNA genes possess characteristics that allow pathogen identification 7 Molecular Detection and Identification of Fusarium oxysporum 139 (Ward 1994; Appel and Gordon 1995; Waalwijk et al. 1996; Edel et al. 2000; Bao et al. 2002; Singh et al. 2006). Bateman et al. (1996) used PCR-RFLP of a PCR product consisting of ITS1, 5.8S and ITS2 ribosomal DNAs, and eight restriction enzymes to distinguish 18 Fusarium haplotypes, while Edel et al. (1997) analyzed further into the 50 end of the 28S rDNA gene to distinguish five Fusarium haplotypes. However, neither of these methods could distinguish among F. crookwellense, F. culmorum, and F. graminearum, indicating that these formae speciales might be more closely related. Indeed, Schilling et al. (1996) later found that the DNA sequence of ITS1 region from F. culmorum and F. graminearum was identical. Additionally, species-specific primers could not be designed due to minor differences in the ITS2 region of the two Fusarium species. Mishra et al. (2003) developed a fluorescent marker-based PCR assay for rapid and reliable identification of five toxigenic and pathogenic Fusarium species viz. F. oxysporum, F. avenaceum, F. culmorum, F. equiseti, and F. sambucinum. The method was based on PCR amplification of species-specific DNA fragments using fluorescent oligonucleotide primers designed from ITS region of rDNA. Similarly, Abd-Elsalam (2003) developed taxon-selective primers using ITS sequences for quick identification of the Fusarium genus, while Abd-Elsalam et al. (2006) identified F. oxysporum f.sp. vasinfectum (Fov) using specific primers based on the 16S and 23S rRNA genes. Based on differences in ITS sequences of Fusarium and Mycosphaerella spp., Zhang et al. (2005) developed species-specific PCR assays for rapid and accurate detection of F. oxysporum f.sp. niveum and Mycosphaerella melonis from diseased watermelon plants and infested soil. They also developed real-time quantitative PCR assays to detect and monitor the pathogens directly in soil samples. Zambounis et al. (2007) used PCR-RFLP and realtime PCR for detection and quantification of Australian isolates of F. oxysporum f.sp. vasinfectum. PCR-RFLP based on the rDNA-IGS region distinguished these isolates from other formae speciales of F. oxysporum. Further, they identified single-nucleotide polymorphisms (SNPs) in the 50 portion of the IGS region and developed two specific real-time PCR assays based on these SNPs for absolute quantification of genomic DNA from the isolates obtained from infected cotton tissues as well as soil samples. Similarly, three Fusarium species from Dendrobium were characterized by Latiffah et al. (2009) using PCR-RFLP of ITS in 5.8S rRNA region. They found that isolates from the same species produced similar PCR-RFLP patterns and UPGMA cluster analysis of the data clearly grouped F. oxysporum, F. proliferatum, and F. solani into separate clusters. Likewise, Gurjar et al. (2009) differentiated F. oxysporum f.sp. ciceri race 3 from the races 1, 2, and 4 based on the polymorphisms obtained with ITS-RFLP and ISSR approaches. 7.2.2.2 Transposons Mouyna et al. (1996) analyzed the South American populations of F. oxysporum f.sp. elaeidis (an oil palm pathogen) and found that they had the palm transposon. They also showed that the palm transposon was present in all the pathogenic 140 R. Saikia and N. Kadoo isolates, but was absent in all the nonpathogenic isolates, indicating that the pathogenic populations may be marked by the transposon. Fernandez et al. (1998) designed specific primers for detection of F. oxysporum f.sp. albedinis (the date palm pathogen), based on the sequences of transposable element Fot1. They analyzed a large number of Fusarium isolates, including 286 F. oxysporum f.sp. albedinis isolates, 17 other formae speciales, nonpathogenic F. oxysporum isolates, and eight other Fusarium species and the specific primer amplified a 400-bp fragment only in F. oxysporum f.sp. albedinis. A diagnostic PCR assay to detect pathogenic F. oxysporum races causing wilt in carnation was developed by Chiocchetti et al. (1999). This strategy was based on the genetic characterization of strains using different transposons and cloning and sequencing the regions flanking the insertion sites of these elements, followed by construction of race-specific primers for quick pathogen identification. Using a similar approach, Pasquali et al. (2007) developed inter-retrotransposon sequence characterized amplified regions (IRSCAR) technique to differentiate F. oxysporum f.sp. lactucae race 1 isolates from other F. oxysporum and F. oxysporum f.sp. lactucae isolates. Interestingly, the robust RAPD marker-based assay developed by Lievens et al. (2007) to specifically detect and identify the economically important cucumber pathogen F. oxysporum f.sp. radicis-cucumerinum showed strong similarity with Folyt1, a transposable element identified in the tomato wilt pathogen F. oxysporum f.sp. lycopersici. 7.2.2.3 Other Genes Mule et al. (2004) developed PCR assays for rapid identification of F. oxysporum and F. proliferatum in asparagus plants based on the calmodulin gene sequences, while Hirano and Arie (2006) designed specific primer sets based on nucleotide differences in endo-polygalacturonase (pg1) and exo-polygalacturonase (pgx4) genes from F. oxysporum f.sp. lycopersici and radicis-lycopersici, infecting tomato. A combination of amplifications from four primer sets allowed effective differentiation of the isolates into formae speciales and races. A PCR-RFLP technique based on TEF-1a gene sequences was designed by Bogale et al. (2007) to distinguish Fusarium redolens and three formae speciales of F. oxysporum. van der Does et al. (2008) found that, despite their polyphyletic origin, the F. oxysporum f.sp. lycopersici isolates contained an identical genomic region of at least 8 kb that was absent in other formae speciales as well as nonpathogenic isolates, and comprised the genes SIX1, SIX2, and SHH1. They further found that SIX3, which lies elsewhere on the same chromosome, was also unique to F. oxysporum f.sp. lycopersici isolates. Recently, five different approaches viz. gene-specific markers, sequence analysis of TEF-1a, ITS-RFLP, ISSR, and AFLP were used distinguish four F. oxysporum f.sp. ciceri (Foc) races, infecting chickpea (Gurjar et al. 2009). Unique AFLP patterns identified the races 1 and 2, while race 4 was distinguished from other races by the absence of amplification product of xylanase-3 gene in this race. The Foc race 3 was differentiated from races 1, 2, and 4 based on the polymorphisms 7 Molecular Detection and Identification of Fusarium oxysporum 141 obtained with ITS-RFLP and ISSR approaches as well as amplification profiles of Hop78 transposon, cutinase, and desaturase genes. However, phylogenetic analysis of TEF-1a data from the four races revealed that race 3 was actually Fusarium proliferatum and not F. oxysporum as has been considered till now (Gurjar et al. 2009). 7.2.2.4 Multiplex PCR Multiplex PCR allows simultaneous and sensitive detection of different DNA or RNA targets in a single reaction. It can therefore, be designed to determine the presence of more than one pathogen in plant material by selectively amplifying specific sequences in two or more of them, or to detect related pathogens on multiple hosts (Louws et al. 1999). Simultaneous identification of several plant pathogens using multiplex PCR has been reported by Hamelin et al. (1996) and de Haan et al. (2000). Demeke et al. (2005) developed a species-specific PCR assay for identification of nine Fusarium species viz. avenaceum, acuminatum, crookwellense, culmorum, equiseti, graminearum, poae, pseudograminearum, and sporotrichioides in pure mycelial culture. Later, they could also simultaneously and accurately identify F. culmorum, F. graminearum, and F. sporotrichioides using multiplex PCR. If such specific primers are developed for common F. oxysporum formae speciales or physiological races, it would greatly simplify their multiplexed detection and identification for timely disease control. However, development of an efficient multiplex PCR requires optimization of reaction conditions in order to discriminate several amplicons per reaction (Elnifro et al. 2000). 7.2.3 Limitations of PCR-Based Techniques Although PCR-based techniques are rapid, highly sensitive, and specific, they might suffer from robustness (van der Wolf et al. 2001). The failure of PCR amplification to correctly diagnose infected and noninfected plant material has been reported in different comparative assays. Carry-over contamination of amplicons could be responsible for false-positive results, while the presence of inhibitor components in sample extracts is the main reason for false negatives. Similarly, PCR based techniques (except reverse transcriptase-PCR) can amplify the target DNA sequences from both active and nonactive or dead pathogen cells/spores (Malorny et al. 2003). Therefore, PCR might yield false positive results in some cases. Another important limitation of PCR-based identification assays is that the technique is not immediately quantitative. Although it is comparatively easy to quantify the amount of a PCR product produced as a result of a successful PCR amplification, it is difficult to estimate the amount of target DNA initially present at the start of the reaction. This is because the reaction rate is exponential; as a result, slight variations in the amplification procedure can generate different amounts of 142 R. Saikia and N. Kadoo final product from the same amount of starting material. Although, target DNA can be quantified using competitive PCR (Nicholson et al. 1998), this method is labor intensive. However, many of these limitations could be overcome by using modern techniques like real-time PCR and microarrays, which are increasingly being used for routine pathogen identification. 7.3 Recent Techniques for Identification of Fusarium oxysporum Currently, the detection of plant pathogens is a changing, dynamic, and evolving world where established protocols can be modified or optimized only months after having been developed. Accurate and routine pathogen detection requires high levels of specificity, sensitivity, reliability, and speed. In this context, specificity can be defined as the capability to detect the pathogen in the absence of false positives and negatives, while sensitivity relates to the lowest number of pathogens reliably detected per assay or sample (Lopez et al. 2003). In addition, pathogen quantification is also becoming important, since it serves as a basis for establishing damage thresholds at which a pathogen causes disease, and action thresholds that determine when measures should be taken to limit or prevent losses (Lievens et al. 2008). As F. oxysporum is known to survive and remain latent in soil for many years, detection methods of high sensitivity, specificity, and reliability are required. The battery of available techniques and probes for detection of plant pathogens has increased considerably over the last few years. In addition to time benefits, there are great advantages in terms of specificity, sensitivity, and reliability with these techniques, as well as, they allow identification of the pathogen camouflaged by other microorganisms. Some of such modern techniques currently used in identification of plant pathogens are discussed below. 7.3.1 Real-Time PCR The real-time PCR technology provides escalating opportunities to identify phytopathogenic fungi and has been used in several studies for detection and identification of various formae speciales of F. oxysporum (Table 7.2). It can more accurately quantify the extent of pathogen biomass in the host tissue and, with multiplex formats, enables simultaneous detection of different pathogens (Lievens et al. 2003). The main advantage of real-time PCR assay over end-point quantitative PCR is that the amplification products can be monitored in real time as they are accumulated in the exponential phase (Schena et al. 2004), thus allowing precise measurement of fungal DNA content in the reaction. 7 Molecular Detection and Identification of Fusarium oxysporum 143 Table 7.2 Some examples of real time-PCR assays developed for detection and identification of Fusarium oxysporum formae speciales F. oxysporum forma Host plant Chemistry References specialis basilici Ocimum basilicum Taqman Pasquali et al. (2006) chrysanthemi Argyranthemum Taqman Pasquali et al. (2004) frutescens chrysanthemi Chrysanthemum Taqman Pasquali et al. (2004) morifolium cucumerinum Cucumis sativus SYBR Green I Lievens et al. (2007) radicis-cucumerinum Cucumis sativus SYBR Green I Lievens et al. (2007) tracheiphilum Vigna unguiculata Taqman Pasquali et al. (2004) tracheiphilum Glycine max Taqman Pasquali et al. (2004) vasinfectum Gossypium spp. SYBR Green I Abd-Elsalam et al. (2006) niveum Citrullus lanatus SYBR Green I Zhang et al. (2005) Pasquali et al. (2004) developed a real-time PCR assay based on TaqMan chemistry to identify a new group of F. oxysporum f.sp. chrysanthemi isolates highly pathogenic on Paris daisy. They successfully identified infected plants using real-time PCR as early as the fifth day after artificial inoculation, although the plants remained symptomless until the 13th day after inoculation. Zhang et al. (2005) used real-time PCR to identify and quantify F. oxysporum f.sp. niveum and Mycosphaerella melonis pathogens directly from soil samples. Similarly, Abd-Elsalam et al. (2006) used real-time PCR based on the 16S and 23S rRNA genes to detect F. oxysporum f.sp. vasinfectum (Fov) in cotton. The assay detected as low as 200 fg of Fov genomic DNA in infected cotton roots, while no amplification was obtained from other plant structures such as stem and leaf. Lievens et al. (2007) developed a robust RAPD marker-based assay to specifically detect and identify the economically important cucumber pathogens F. oxysporum f.sp. cucumerinum and F. oxysporum f.sp. radicis-cucumerinum. They used the real-time PCR assay to confirm that the selected RAPD markers for F. oxysporum f.sp. cucumerinum and F. oxysporum f.sp. radicis-cucumerinum represented single copy DNA sequences. Likewise, Zambounis et al. (2007) developed two specific real-time PCR based assays based on the SNPs found in the 50 portion of the rDNA-IGS regions for quantification of genomic DNA of Australian isolates of F. oxysporum f.sp. vasinfectum from infected cotton tissues as well as soil samples. However, like all other molecular methods based on DNA amplification, a major drawback of the system is that it is unable to distinguish between viable and dead propagules. Similarly, multiplexing in real-time PCR is limited by the number of different fluorescent dyes available. In addition, the initial and running costs of a real-time PCR system are several times more than a normal PCR system. However, considering the many benefits of the real-time PCR technology compared to normal PCR, the use of real-time PCR is still advantageous. Higgins et al. (2003) developed a portable real-time PCR instrument for performing diagnostic assays directly in the field. Such rapid real-time PCR diagnosis could result in taking appropriate and timely control measures than possible with traditional methods of pathogen 144 R. Saikia and N. Kadoo identification. Therefore, the resulting losses due to diseases as well as the cost of disease management could be greatly minimized. 7.3.2 Microarrays Microarray holds promise for quick and accurate detection of plant pathogens. The potential of microarray technology in the detection and identification of plant pathogens is very high due to multiplex capabilities of the system. The application of microarrays for detection of pathogens in various environments has enabled parallel detection of multiple species in a high throughput format conducive to automation (Small et al. 2001; Loy et al. 2002). For pathogenic fungi, microarray analysis has a great potential to systematically and efficiently identify genes required for infection (Lorenz 2002; Bryant et al. 2004). A Magnaporthe grisea array is now commercially available from Agilent Technologies (http://www. agilent.com/). A molecular detection system based on DNA array technology was developed by Lievens et al. (2003) for rapid and efficient detection of tomato vascular wilt pathogens F. oxysporum f.sp. lycopersici, Verticillium albo-atrum, and V. dahliae. The array was successfully used for sensitive detection of the tomato wilt pathogens from complex substrates like soil, plant tissues, and irrigation water as well as samples collected from tomato growers. Similarly, microarray analysis of F. oxysporum f.sp. vasinfectum genes expressed in planta (McFadden et al. 2006), has revealed pathogenic genes in the cotton pathogen. The expression of this gene was also positively correlated with vascular browning, which is a characteristic symptom of Fusarium wilt infection (McFadden et al. 2006). Guldener et al. (2006) reported the design and validation of the first Affymetrix GeneChip microarray based on the entire genome of Fusarium graminearum. It has been shown to efficiently detect genes from four other closely related species of Fusarium graminearum. As the genomes of some formae speciales of F. oxysporum have already been sequenced by the Broad Institute (http://www.broad.mit.edu/), microarray chips might become available for these and other formae speciales of F. oxysporum in the near future. Generally, one needs to analyze conserved genes when taxonomically comparing phyla, orders, families, or genera. However, less conserved genes must be used when investigating species within a genus or taxonomic levels below the species (Lévesque 2001), such as formae speciales and races. Considering this, Lievens et al. (2007) developed a DNA array containing genus-, species- and forma specialis-specific detector oligonucleotides for the detection and identification of F. oxysporum f.sp. cucumerinum and F. oxysporum f.sp. radicis-cucumerinum. The array utilized the rRNA gene cluster to derive genus- and species-specific oligonucleotides, whereas RAPD-derived SCAR markers were used to specifically identify different formae speciales. Using such approach and taking into account the almost unlimited expanding possibilities of DNA arrays, a comprehensive DNA array for 7 Molecular Detection and Identification of Fusarium oxysporum 145 the identification of all formae speciales (and possibly even races) of F. oxysporum may ultimately be realized (Lievens et al. 2008). 7.3.3 Gene/Genome Sequencing One of the most robust and informative techniques useful in fungal diagnosis is nucleotide sequencing, where DNA sequence variations can be used to design species-specific primers and/or probes. Sequences of the TEF-1a and the mitochondrial small subunit (mtSSU) ribosomal RNA genes have been valuable in distinguishing different species (Baayen et al. 2000, 2001; O’Donnell et al. 2000; Skovgaard et al. 2001). Phylogenetic analysis of TEF-1a data by Gurjar et al. (2009) from four F. oxysporum f.sp. ciceri races revealed that race 3 of the pathogen was actually Fusarium proliferatum and not F. oxysporum as has been considered till now. Similarly, DNA sequences of UTP-ammonia ligase, trichothecene 3-Oacetyltransferase, a putative reductase (O’Donnell et al. 2000), nitrate reductase and phosphate permease (Skovgaard et al. 2001) have also been used successfully to distinguish different Fusarium species. Elucidation of full sequence of the genome of an organism can help in designing the most accurate and sensitive method for its detection and identification. Based on genome analysis, new specific sequences could be used to design microarray chips, detection probes, or PCR primers for different pathogens. Using these, it is possible to identify not only the formae speciales, but also individual races of a pathogen. Genome sequencing efforts are currently in progress for several important species of the Fusarium genus (F. circinatum, F. graminearum, F. oxysporum, F. proliferatum, F. sporotrichioides, and F. verticillioides) (Table 7.3) and the genomes are available at Genomes OnLine Database (GOLD; http://www.genomesonline.org/ index2.htm). In case of F. oxysporum, the genome of f.sp. lycopersici strain FGSC 4286 was sequenced at 6.8X coverage using the whole genome shotgun (WGS) sequencing method and the draft sequence is now available. Automated annotation of the draft sequence predicted over 17,000 protein-coding genes. Further analysis of these genes can elucidate host-pathogen interactions and allow the development of disease management approaches targeting important pathogenicity genes of the pathogens. 7.3.4 DNA Barcoding DNA barcoding holds enormous potential for the rapid identification of organisms at the species level. It is a taxonomic method that uses a short genetic marker in an organism’s DNA to identify it as belonging to a particular species. It is emerging as an important tool for the precise taxonomic identification of a wide range of species and is effective at both identifying existing species and discovering newones. 146 R. Saikia and N. Kadoo Table 7.3 Status of genome sequencing projects of various Fusarium speciesa Type Genome Sequencing centers Sequncing Fusarium species/ strains size depth F. circinatum FSP 34 Genome 50 Mb Inqaba 10x Biotechnologies FABI F. graminearum Genome 36 Mb Ceral Disease N.A. Laboratory F. graminearum Genome 36 Mb Syngenta AG N.A. F. graminearum PH-1 Genome 36 Mb Broad Institute 10x IGGR Genome 17,000 Broad Institute 6.8x F. oxysporum f.sp. lycopersici FGSC ORFs 4286 F. proliferatum Genome N.A. Greenomics N.A. F. sporotrichioides EST N.A. Univ of Oklahoma N.A. F. verticillioides Genome 87 Mb J. Craig Venter N.A. Institute USDA/ ARS F. verticillioides 7600 Genome 46 Mb Broad Institute 8x Syngenta a Status as of August 31, 2009; N.A.: Not available Sequencing status Complete Incomplete Incomplete Complete and Published Incomplete Incomplete Incomplete Complete Incomplete A 648-bp region of the mitochondrial cytochrome c oxidase subunit I (COI) gene was initially proposed as a potential “DNA barcode” (Hebert et al. 2003). Using this new standard, databases are being developed to facilitate rapid and accurate identification of plant pathogens in general (Plant Pathogen Barcode, PPB; http:// www.plantpathogenbarcode.org/) and fungi in particular (All Fungi Barcoding, http://www.allfungi.com/). Seifert et al. (2007) evaluated suitability of the COI gene in fungi by analyzing 370 strains from 58 species of Penicillium subgenus Penicillium and 12 allied species and found that the gene could be successfully used for fungal barcoding. Strains from 38 out of 58 species formed cohesive assemblages with distinct COI sequences, and all cases of sequence sharing involved known species complexes. However, there are reports that the COI gene does not work well for most true fungi and some researchers feel that the most appropriate gene for DNA barcoding of true fungi is the ITS region of the nuclear ribosomal DNA (http://www.allfungi.com/). Although DNA barcoding holds great promise for species identification, its use in molecular phylogenetics is challenging. Hajibabaei et al. (2006) showed that phylogenetic trees constructed from short DNA barcodes, although approximately reflected accepted phylogenetic relationships, had low statistical support at many of the internal nodes and could seriously misrepresent some of the branching patterns. Hence, Min and Hickey (2007) assessed the effect of varying sequence length of DNA barcodes for the classification of unknown specimens at the species level as well as for phylogenetic reconstruction in fungi. They found that reducing the length of the barcode had a profound effect on the accuracy of resulting phylogenetic trees; however, the short barcode sequences still identified the fungal species accurately. They concluded that the standard short barcode sequences (~600 bp) 7 Molecular Detection and Identification of Fusarium oxysporum 147 were suitable for species identification, but not for inferring accurate phylogenetic relationships among the fungi. Hence, it is possible that the standard DNA barcoding might accurately distinguish different Fusarium species; however, longer barcodes would be necessary to precisely identify different formae speciales and races of the F. oxysporum species complex. 7.4 7.4.1 Emerging Technologies for Pathogen Identification Next-Generation Sequencing The recently-developed “Next-Generation” sequencing platforms, such as 454 (Roche), Solexa (Illumina), and SOLiD (ABI), allow researchers to obtain several million bp of sequences affordably in a single run in an unbiased manner. Among these sequencing platforms, the 454 GS FLX instrument currently has the ability to sequence 400–600 million bp per run (with 400–500 bp individual reads) using the Titanium series reagents (http://www.454.com/). Due to its high accuracy, low cost, and long reads compared to the Solexa and SOLiD systems, many researchers have migrated toward the 454 sequencing platform for a variety of genome projects. As these instruments have the potential of sequencing several microbial genomes in a single run, it is very likely that the genomes of economically important plant pathogens, including various Fusarium species, will be shortly available. Indeed, genome sequencing projects of several Fusarium species are already in progress (Table 7.3). Based on the analysis of these genomes, specific oligonucleotide sequences could be used to design microarray chips, detection probes, or PCR primers for high-throughput or multiplexed detection and identification of different F. oxysporum strains. If the genomes of important F. oxysporum formae speciales, and individual races become available, the pathogenic isolates could be detected specifically even if camouflaged by other organisms. 7.4.2 Single-Nucleotide Polymorphisms Detection and characterization of SNPs is also one of the promising post-genomics research tools for pathogen identification. This new technology is pushing pathogen identification to its ultimate limit-the single base pair difference. It is presumed that many plant pathogenic races or formae speciales differ from their closest relatives by only a few bases in different genes. The next-generation sequencing platforms can rapidly carryout deep sequencing of microbial genomes, enabling quick discovery of SNPs in different pathogenic strains of the microbial species. This will enable designing forma specialis or race-specific cleaved amplified polymorphic sequence (CAPS) or derived cleaved amplified polymorphic sequence (dCAPS) 148 R. Saikia and N. Kadoo markers for PCR-based identification. CAPS markers result from differential restriction digestion of gene/allele specific PCR products based on the loss or gain of restriction enzyme recognition sites due to the presence of SNPs or insertion/deletion mutations. While in dCAPS analysis, a restriction enzyme recognition site that includes the SNP is introduced into the PCR product by a primer containing one or more mismatches to template DNA (Neff et al. 1998). Zambounis et al. (2007) discovered SNPs in a portion of the IGS region of rDNA flanking the 50 end and developed specific real-time PCR-based assays for absolute quantification of genomic DNA from the Australian Fusarium oxysporum f.sp. vasinfectum isolates obtained from infected cotton tissues as well as soil samples. 7.4.3 Metagenomics Another promising approach to large scale detection of microbes from diverse samples is metagenomics. It is the study of genomic content of microbial organisms directly in their natural environments, bypassing the need for isolation and culturing of individual species (Chen and Pachter 2005). Hence, metagenomics enables studies of organisms that are not culturable as well as studies of organisms in their natural environment. Using the metagenomics approach, these new sequencing technologies enable researchers to quickly and affordably identify the organisms present in a complex sample (such as soil, irrigation water, or plant tissues) without any prior knowledge. Such metagenomics approach to pathogen identification should facilitate quick identification and quantification of a range of pathogens present in the sample and enable undertaking appropriate disease control strategies well before the pathogen populations reach damage thresholds. The 454 GS FLX System is very suitable for metagenomics as the system’s long reads help in accurate identification of pathogenic strains present in the sample. Researchers often use the platform for counting gene tags to analyze the relative abundance of different microbial species in different samples. 7.5 Potential Limitations of Molecular Identification Techniques Currently, PCR, real-time PCR, and microarrays are the methods of choice for rapid and accurate detection of plant pathogens. However, a major problem of PCR based detection is that PCR could amplify DNA from both active and non-active or dead pathogen cells/spores. Therefore, it may yield false positive results in some cases. Similarly, false negatives can be attributed in standard PCR protocols due to the presence of compounds that inhibit the polymerases, degradation of the target DNA sequence, or reagent problems (Louws et al. 1999). Likewise, although microarray 7 Molecular Detection and Identification of Fusarium oxysporum 149 is the most suitable technique for multiplexed detection of many isolates of F. oxysporum and other pathogens in a single assay, currently microarrays are expensive for routine application. Moreover, additional work is needed to address the challenges of working with environmental samples where contaminants may interfere with DNA hybridization and affect the performance of microarrays. Similarly, the lack of adequate sequence information can hamper the development of reliable molecular diagnostic assays. Moreover, techniques like DNA barcoding are presently unable to differentiate pathogenic strains from non-pathogenic ones that belong to the same microbial species. Hence, if no molecular markers are available to distinguish the pathogenic subspecies, pathogenicity test is the only way to determine whether or not a given isolate is pathogenic on a specific crop or variety. Although technically feasible and potentially invaluable, large sequencing studies still face significant challenges. Foremost among the challenges is analyzing the tremendous amounts of data generated (Nelson 2003). It is relatively easy to characterize genes and genome of a well-studied and easily cultivated microbe; however, it would be a daunting task to understand the genomics of unknown or uncultured microbes or whole environmental genomes revealed by metagenomics approaches. For example, Tringe et al. (2005) could assemble as many as 150,000 sequence reads into contiguous sequences spanning only 1% of a soil metagenome. They estimated that 2–5 billion bp of sequence would be needed to completely cover the metagenome of a Minnesota soil. Similarly, in mixed microbial communities like agricultural soils, it will be difficult to separate, assemble, and annotate the genomes of a range of soil microflora. In addition, incorrect assembly of contiguous genomes and the formation of chimeric inserts can create problems in interpreting the data (Schloss and Handelsman 2005). 7.6 Conclusions and Future Prospects Genomics research is generating fast-growing databases that can be used to design molecular assays for simultaneous detection of a large number of pathogens, coupled with novel platforms having unprecedented capabilities for multiplexing, high throughput, and portability. As these new technologies gain wide acceptance, routine detection, identification, and monitoring of plant pathogens should become more common in plant pathology. Microarray chips are now being fabricated with oligonucleotides that are either synthesized directly on a solid surface or are microspotted. Similarly, the next generation sequencing technologies like 454 and SOLiD can sequence several microbial genomes in a single run. If the complete DNA sequence of plant pathogens is revealed, oligonucleotides specific for a pathogen can be designed and a single high-density microarray chip could accommodate oligonucleotides specific to a large number of pathogens. In the next few years, complete genome sequences of many pathogenic strains of F. oxysporum are likely to become available and these will help to design 150 R. Saikia and N. Kadoo PCR primers or probes very specific to the pathogen strains, enabling accurate identification of the strains even if camouflaged by other pathogens. For example, if each microarray chip contained oligonucleotides specific to each of the known formae speciales and races of F oxysporum, it would be possible to have multiplex detection of all these pathogens in one experiment even from complex substrates like soil, plant tissues, and irrigation water. With such high-throughput technologies, integration of more strains into the detection systems of F. oxysporum should become possible and identification of pathogens is likely to become an easier task. However, these should be observed as management tools, to be used in combination with the knowledge of the crop and understanding of the biology of different formae speciales and ecology of the diseases. In this respect, the increasing availability of full-genome sequences of many plant pathogens including formae speciales of F. oxysporum is a welcome development. With the availability of affordable and portable real-time PCR instruments (Higgins et al. 2003) and simpler protocols, molecular-based diagnosis of crop diseases is becoming a field reality. Routine diagnosis of many crop diseases is now possible in one day or less using recent innovative technologies. This, coupled with high throughput that reduces the cost per sample, should make these assays more attractive for use in crop protection. A combination of DNA microarrays with other genomic methods will certainly accelerate the efforts to characterize the function of unknown stretches of fungal genomes. The resulting database will allow complete analysis of developmental processes that are characteristics of the fungus, including the molecular nature of pathogenicity. However, new molecular detection technologies that are portable, robust, sensitive, and cost effective need to be developed for routine identification of plant pathogens directly in the field to undertake appropriate disease control measures as quickly as possible. References Abd-Elsalam KA (2003) Non-gel based techniques for plant pathogen genotyping. Acta Microbiol Polon 52:329–341 Abd-Elsalam KA, Khalil MS, Aly AA, Asran-Amal A (2002a) Genetic diversity among Fusarium oxysporum f.sp. vasinfectum isolates revealed by UP–PCR and AFLP markers. Phytopathologia Medit 41:1–7 Abd-Elsalam KA, Khalil MS, Aly AA, Asran-Amal A (2002b) Population analysis of Fusarium spp. Phytomedizin 3:18–19 Abd-Elsalam KA, Omar MR, Migheli Q, Nirenberg HI (2004) Genetic characterization of Fusarium oxysporum f.sp. vasinfectum isolates by random amplification of polymorphic DNA (RAPD) and amplified fragment length polymorphism (AFLP). J Plant Dis Protect 111:534–544 Abd-Elsalam KA, Asran-Amal A, Schnieder F, Migheli Q, Verreet JA (2006) Molecular detection of Fusarium oxysporum f.sp. vasinfectum in cotton roots by PCR and real-time PCR assay. J Plant Dis Protect 113:14–19 Abdel-Satar MA, Khalil MS, Mohmed IN, Abd-Elsalam KA, Verreet JA (2003) Molecular phylogeny of Fusarium species by AFLP fingerprint. Afr J Biotechnol 2:51–55 7 Molecular Detection and Identification of Fusarium oxysporum 151 Abo K, Klein KK, Edel-Hermann V, Gautheron N, Traore D, Steinberg C (2005) High genetic diversity among strains of Fusarium oxysporum f.sp. vasinfectum from cotton in Ivory Coast. Phytopathology 95:1391–1396 Alves-Santos FM, Benito EP, Eslava AP, Diaz-Minguez JM (1999) Genetic diversity of Fusarium oxysporum strains from common bean fields in Spain. Appl Environ Microbiol 65:3335–3340 Alves-Santos FM, Garcı́a-Sánchez RB, MA EAP, Dı́az-Mı́nguez JM (2002) A DNA-based procedure for in planta detection of Fusarium oxysporum f.sp. phaseoli. Phytopathology 92:237–244 Appel DJ, Gordon TR (1995) Intraspecific variation within populations of Fusarium oxysporum based on RFLP analysis of the intergenic spacer region of the rDNA. Exp Mycol 19:120–128 Armstrong GM, Armstrong JK (1981) Formae specialis and races of Fusarium oxysporum causing wilt diseases. Fusarium: diseases, biology and taxonomy. Pennsylvania State University Press, University Park (US), pp 391–399 Assigbetse KB, Fernandez D, Dubois MP, Geiger JP (1994) Differentiation of Fusarium oxysporum f.sp. vasinfectum races on cotton by random amplified polymorphic DNA (RAPD) analysis. Phytopathology 84:622–626 Attitalla IH, Fatehi J, Levenfors J, Brishammar S (2004) A rapid molecular method for differentiating two special forms (lycopersici and radicis-lycopersici) of Fusarium oxysporum. Mycol Res 108:787–794 Baayen RP, van Dreven F, Krijger MC, Waalwijk C (1997) Genetic diversity in Fusarium oxysporum f.sp. dianthi and Fusarium redolens f.sp. dianthi. Eur J Plant Pathol 103:395–408 Baayen RP, Forch MG, Waalwijk C, Bonants PJM, Loffler HJM, Roebroeck EJA (1998) Pathogenic, genetic, and molecular characterization of Fusarium oxysporum f.sp. lilii. Eur J Plant Pathol 104:887–894 Baayen RP, O’Donnell K, Bonants PJM, Cigelnik E, Kroon LPNM, Roebroeck EJA, Waalwijk C (2000) Gene genealogies and AFLP analyses within the Fusarium oxysporum complex identify monopbyletic and non-monophyletic Formae specialis causing wilt and rot disease. Phytopathology 90:891–900 Baayen RP, O’Donnell K, Breeuwsma S, Geiser DM, Waalwijk C (2001) Molecular relationships of fungi within the Fusarium redolens-F hostae clade. Phytopathology 91:1037–1044 Bao JR, Fravel DR, O’Neill NR, Lazarovits G, van Berkum P (2002) Genetic analysis of pathogenic and non-pathogenic Fusarium oxysporum from tomato plant. Can J Bot 80: 271–279 Barve MP, Haware MP, Sainani MN, Ranjekar PK, Gupta VS (2001) Potential of microsatellites to distinguish four races of Fusarium oxysporum f.sp. ciceri prevalent in India. Theor Appl Genet 102:138–147 Bateman GL, Kwasna H, Ward E (1996) Relationship among Fusarium spp estimated by comparing restriction fragment length polymorphisms in polymerase chain reaction-amplified nuclear rDNA. Can J Microbiol 42:1232–1240 Bayraktar H, Dolar FS, Maden S (2008) Use of RAPD and ISSR markers in detection of genetic variation and population structure among Fusarium oxysporum f.sp. ciceris isolates on chickpea in Turkey. J Phytopathol 156:146–154 Beckman CH, Roberts EM (1995) On the nature and genetic basis for resistance and tolerance to wilt diseases of plants. Adv Bot Res 21:35–77 Bentley S, Pegg KG, Dale JL (1994) Optimization of RAPD-fingerprinting 1 to analyze genetic variation within populations of Fusarium oxysporum f.sp. cubense. J Phytopathol 142:64–78 Bentley S, Pegg KG, Dale JL (1995) Genetic variation among a world-wide collection of isolates of Fusarium oxysporum f.sp. cubense analyzed by RAPD–PCR fingerprinting. Mycol Res 11:1378–1384 Bodker L, Lewis BG, Coddington A (1993) The occurrence of a new genetic variant of Fusarium oxysporum f.sp. pisi. Plant Pathol 42:833–838 152 R. Saikia and N. Kadoo Boehm EWA, Ploetz RC, Kistler HC (1994) Statistical analysis of electrophoretic karyotype variation among vegetative compatibility groups of Fusarium oxysporum f.sp. cubense. Mol Plant Microbe Interact 7:196–207 Bogale M, Wingfield BD, Wingfield MJ, Steenkamp ET (2005) Simple sequence repeat markers for species in the Fusarium oxysporum complex. Mol Ecol Notes 5:622–624 Bogale M, Wingfield BD, Wingfield MJ, Steenkamp ET (2006) Characterization of Fusarium oxysporum isolates from Ethiopia using AFLP, SSR and DNA sequence analyses. Fungal Div 23:51–66 Bogale M, Wingfield BD, Wingfield MJ, Steenkamp ET (2007) Species-specific primers for Fusarium redolens and a PCR-RFLP technique to distinguish among three clades of Fusarium oxysporum. FEMS Microbiol Lett 271:27–32 Bosland PW, Williams PH (1987) An evaluation of Fusarium oxysporum from crucifers based on pathogenicity, isozyme polymorphism, vegetative compatibility, and geographic origin. Can J Bot 65:2067–2073 Bryant P, Venter A, Robins-Browne D, Curtis R (2004) Chips with everything: DNA microarrays in infectious diseases. Lancet Infect Dis 4:100–111 Burgess LW (1981) General ecology of the Fusaria. In: Nelson PE, Toussoun TA, Cook RJ (eds) Fusarium: diseases, biology and taxonomy. Pennsylvania State University Press, University Park, Pennsylvania, USA, pp 225–235 Chen K, Pachter L (2005) Bioinformatics for whole-genome shotgun sequencing of microbial communities. PLoS Comp Biol 1:24 Chiocchetti A (2001) PCR Detection of Fusarium oxysporum f. sp. basilici on Basil. Plant Dis 85:607–611 Chiocchetti A, Ghignone S, Minuto A, Gullino ML, Garibaldi A, Migheli Q (1999) Identification of Fusarium oxysporum f.sp. basilici isolated from soil, basil seed, and plants by RAPD analysis. Plant Dis 83:576–581 Cramer RA, Byrne PF, Brick MA, Panella L, Wickliffe E, Schwartz HF (2003) Characterization of Fusarium oxysporum isolates from common bean and sugar beet using pathogenicity assays and random amplified polymorphic DNA markers. J Phytopathol 151:352–360 de Haan LAM, Numansen A, Roebroeck EJA, van Doorn J (2000) PCR detection of Fusarium oxysporum f.sp. gladioli race 1, causal agent of Gladiolus yellows disease, from infected corms. Plant Pathol 49:89–100 Demeke T, Clear RM, Patrick SK, Gaba D (2005) Species-specific PCR based assays for the detection of Fusarium species and a comparison with the whole seed agar plate method and trichothecene analysis. Intl J Food Microbiol 103:271–284 DeScenzo RA, Engel SR, Gomez G, Jackson EL, Munkvold GP, Weller J, Irelan NA (1999) Genetic analysis of Eutypa strains from California supports the presence of two pathogenic species. Phytopathology 89:884–893 Edel V, Steinberg C, Gautheron N, Alabouvette C (1997) Evaluation of restriction analysis of polymerase chain reaction (PCR)- amplified ribosomal DNA for the identification of Fusarium species. Mycol Res 101:179–187 Edel V, Steinberg C, Gautheron N, Alabouvette C (2000) Ribosomal DNA-targeted oligonucleotide probe and PCR assay specific for Fusarium oxysporum. Mycol Res 104:518–526 Edwards SG, O’callaghan J, Dobson ADW (2002) PCR-based detection and quantification of mycotoxigenic fungi. Mycol Res 106:1005–1025 Elias KS, Schneider RW (1991) Vegetative compatibility groups in Fusarium oxysporum f.sp. lycopersici. Phytopathology 81:159–162 Elias KS, Schneider RW (1992) Genetic diversity within and among races and vegetative compatibility groups of Fusarium oxysporum f.sp. lycopersici as determined by isozyme analysis. Phytopathology 82:1421–1427 Elias KS, Zamir D, Lichtman-Pleban T, Katan T (1993) Population: structure of Fusarium oxysporum f.sp. lycopersici: restriction fragment length polymorphisms provide genetic evidence that vegetative compatibility group is an indicator of evolutionary origin. Mol Plant Microbe Interact 6:565–572 7 Molecular Detection and Identification of Fusarium oxysporum 153 Elnifro EM, Ashshi AM, Cooper RJ, Klapper PE (2000) Multiplex PCR: optimization and application in diagnostic virology. Clin Microbiol Rev 13:559–570 Fernandez D, Tantaoui A (1994) Random amplified polymorphic DNA (RAPD) analysis: a tool for rapid characterization of Fusarium oxysporum f.sp. albedinis isolates. Phytopathol Mediterr 33:223–229 Fernandez D, Assigbetse KB, Dubois MP, Geiger JP (1994) Molecular characterization of races and vegetative compatibility groups in Fusarium oxysporum f.sp. vasinfectum. Appl Environ Microbiol 60:4039–4046 Fernandez D, Ouinten M, Tantaoui A, Lourd M, Geiger JP (1995) Population genetic structure of Fusarium oxysporum f.sp. albedinis (Abstr). Fungal Genet Newsl 42A:34 Fernandez D, Ouinten M, Tantaoui A, Geiger JP, Daboussi MJ, Langin T (1998) Fot1 insertions in the Fusarium oxysporum f. sp. albedinis genome provide diagnostic PCR targets for detection of the date palm pathogen. Appl Environ Microbiol 64:633–636 Flood J, Whitehead DS, Cooper RM (1992) Vegetative compatibility and DNA polymorphisms in Fusarium oxysporum f.sp. elaeidis and their relationship to isolate virulence and origin. Physiol Mol Plant Pathol 41:201–215 Garcia-Mas J, Oliver M, G’omez-Paniagua H, de Vicente MC (2000) Comparing AFLP, RAPD and AFLP markers for measuring genetic diversity in melon. Theor Appl Genet 101:860–864 Gerlach KS, Bentley S, Moore NY, Pegg KG, Aitken AB (2000) Characterisation of Australian isolates of Fusarium oxysporum f.sp. cubense by DNA fingerprinting analysis. Aus J Agri Res 51:945–953 Gonzalez M, Rodriguez MEZ, Jacabo JL, Hernandez F, Acosta J, Martinez O, Simpson J (1998) Characterization of Mexican isolates of Colletotrichum lindemuthianum by using differential cultivars and molecular markers. Phytopathology 88:292–299 Grajal-Martin MJ, Simon CJ, Muehlbauer FJ (1993) Use of random amplified polymorphic DNA (RAPD) to characterize race 2 of Fusarium oxysporum f.sp. pisi. Phytopathology 83:612–614 Groenewald S, Berg NVD, Marasas WFO, Viljoen A (2006) The application of high-throughput AFLP’s in assessing genetic diversity in Fusarium oxysporum f.sp. cubense. Mycol Res 110:297–305 Guldener U, Seong KU, Boddu J, Cho S, Trail F, Xu JR, Adam G, Mewes HW, Muehlbauer FJ, Kistler HC (2006) Development of a Fusarium graminearum Affymetrix GeneChip for profiling fungal gene expression in vitro and in planta. Fungal Gen Biol 43:316–325 Gurjar G, Barve M, Giri A, Gupta V (2009) Identification of Indian pathogenic races of Fusarium oxysporum f.sp. ciceris with gene specific, ITS and random markers. Mycologia 101:480–491 Hajibabaei M, Janzen DH, Burns JM, Hallwachs W, Hebert PDN (2006) DNA barcodes distinguish species of tropical Lepidoptera. Proc Natl Acad Sci USA 103:968–971 Hamelin RC, Berube P, Gignac M, Bourassa M (1996) Identification of root rot fungi in nursery seedlings by nested multiplex PCR. Appl Environ Microbiol 62:4026–4031 Hebert PDN, Cywinska A, Ball SL, de Waard JR (2003) Biological identifications through DNA barcodes. Proc R Soc Lond B Biol Sci 270:313–322 Higgins JA, Nasarabadi S, Karns JS, Shelton DR, Cooper M, Gbakima A, Koopman RP (2003) A handheld real time thermal cycler for bacterial pathogen detection. Biosens Bioelectr 18: 1115–1123 Hirano Y, Arie T (2006) PCR-based differentiation of Fusarium oxysporum f.sp. lycopersici and radicis-lycopersici and races of F oxysporum f.sp. lycopersici. J Genet Plant Pathol 72: 273–283 Hirota N, Hashiba T, Yoshida H, Kikumoto T, Ehara Y (1992) Detection and properties of plasmid-like DNA in isolates from twentythree formae specialis of Fusarium oxysporum. Ann Phytopathol Soc Jpn 58:386–392 Honnareddy N, Dubey SC (2006) Pathogenic and molecular characterization of Indian isolates of Fusarium oxysporum f.sp. ciceris causing chickpea wilt. Curr Sci 91:661–666 Jacobson DJ, Gordon TR (1990) Variability of mitochondrial DNA as an indicator of relationships between populations of Fusarium oxysporum f. sp. melonis. Mycol Res 94:734–744 154 R. Saikia and N. Kadoo Jimenez-Gasco MM, Jimenez-Diaz RM (2003) Development of a specific polymerase chain reaction-based assay for the identification of Fusarium oxysporum f.sp. ciceris and its pathogenic races 0, IA, 5, and 6. Phytopathology 93:200–209 Jimenez-Gasco MD, Perz-Artes E, Jimenez-Diaz RM (2001) Identification of pathogenic races 0, 1B/C, 5, and 6 of Fusarium oxysporum f.sp. ciceris with random amplified polymorphic DNA (RAPD). Eur J Plant Pathol 107:237–248 Katan T (1999) Current status of vegetative compatibility groups in Fusarium oxysporum. Phytoparasitica 27:51–64 Katan T, Di Primo P (1999) Current status of vegetative compatibility groups in Fusarium oxysporum. Phytoparasitica 27:273–277 Katan T, Zamir D, Sarfatti M, Katan J (1991) Vegetative compatibility groups and subgroups in Fusarium oxysporum f.sp. radicis-lycopersici. Phytopathology 81:255–262 Kelly A, Alcala-Jimenez AR, Bainbridge BW, Heale JB, Perez-Artes E, Jimenez-Diaz RM (1994) Use of genetic fingerprinting and random amplified polymorphic DNA to characterize pathotypes of Fusarium oxysporum f.sp. ciceris infecting chickpea. Phytopathology 84:1293–1298 Kelly AG, Bainbridge BW, Heale JB, Perez-Artes E, Jimenez-Diaz RM (1998) In planta-polymerase-chain-reaction detection of the wilt-inducing pathotype of Fusarium oxysporum f.sp. ciceris in chickpea (Cicer arietinum L). Physiol Mol Plant Pathol 52:397–409 Kim DH, Martyn RD, Magill CW (1993) Chromosomal polymorphism in Fusarium oxysporum f.sp. niveum. Phytopathology 83:1209–1216 Kiprop EK, Baudoin JP, Mwang’ombe AN, Kimani PM, Mergeai G, Maquet A (2002) Characterization of Kenyan isolates of Fusarium udum from pigeonpea [Cajanus cajan (L) Mill sp] by cultural characteristics, aggressiveness and AFLP analysis. J Phytopathol 150:517–527 Kistler HC (1997) Genetic diversity in the plant-pathogenic fungus Fusarium oxysporum. Phytopathology 87:474–479 Kistler HC, Benny U (1989) The mitochondrial genome of Fusarium oxysporum. Plasmid 22:86–89 Kistler HC, Bosland PW, Benny U, Leong S, Williams PH (1987) Relatedness of strains of Fusarium oxysporum from crucifers measured by examination of mitochondrial and ribosomal DNA. Phytopathology 77:1289–1293 Kistler HC, Momol EA, Benny U (1991) Repetitive genomic sequences for determining relatedness among strains of Fusarium oxysporum. Phytopathology 81:331–336 Koenig RL, Kistler HC, Ploetz RC (1993) Restriction fragment length polymorphism analysis of Fusarium oxysporum f.sp. cubense. In: 6th Int Congr Plant Pathol. National Research Council of Canada, Ottawa, pp 158 Latiffah Z, Nur-Hayati MZ, Baharuddin S, Maziah Z (2009) Identification and Pathogenicity of Fusarium species associated with root rot and stem rot of Dendrobium. Asian J Plant Pathol 3:14–21 Leslie JF, Zeller KA, Lamprecht SC, Rheeder JP, Marasas WFO (2005) Toxicity, pathogenicity, and genetic differentiation of five species of Fusarium from sorghum and millet. Phytopathology 95:275–283 Lévesque CA (2001) Molecular methods for detection of plant pathogens – what is the future. Can J Plant Pathol 24:333–336 Lievens B, Brouwer M, Vanachter ACRC, Lévesque CA, Cammue BPA, Thomma BPHJ (2003) Design and development of a DNA array for rapid detection and identification of multiple tomato vascular wilt pathogens. FEMS Microbiol Lett 223:113–122 Lievens B, Claes L, Vakalounakis DJ, Vanachter ACRC, Thomma BPHJ (2007) A robust identification and detection assay to discriminate the cucumber pathogens Fusarium oxysporum f.sp. cucumerinum and f.sp. radicis-cucumerinum. Environ Microbiol 9:2145–2161 Lievens B, Rep M, Thomma BPHJ (2008) Recent developments in the molecular discrimination of formae speciales of Fusarium oxysporum. Pest Manag Sci 64:781–788 Lopez MM, Bertolini E, Olmos A, Caruso P, Gorris MT, Llop P, Penyalver R, Cambra M (2003) Innovative tools for detection of plant pathogenic viruses and bacteria. Int Microbiol 6: 233–243 7 Molecular Detection and Identification of Fusarium oxysporum 155 Lorenz MC (2002) Genomic approaches to fungal pathogenicity. Curr Opin Microbiol 5:372–378 Louws FJ, Rademaker JLW, Brujin FJ (1999) The three Ds of PCR-based genomic analysis of phytobacteria: diversity, detection, and disease diagnosis. Ann Rev Phytopathol 37:81–125 Loy A, Lehner A, Lee N, Adamczyk J, Meier H, Ernst J, Schleifer K, Wagner M (2002) Oligonucleotide arrays for 16S rRNA genebased detection of all recognized lineages of sulfate-reducing Prokaryotes in the environment. Appl Environ Microbiol 68:5064–5081 Majer D, Mithen R, Lewis B, Vos P, Oliver RP (1996) The use of AFLP fingerprinting for the detection of genetic variation in fungi. Mycol Res 100:1107–1111 Malorny B, Tassios PT, Radstrom P, Cook N, Wagner M, Hoorfar J (2003) Standardization of diagnostic PCR for the detection of foodborne pathogens. Intl J Food Microbiol 83:39–48 Manicom BQ, Baayen RP (1993) Restriction fragment length polymorphisms in Fusarium oxysporum f.sp. dianthi and other fusaria from Dianthus species. Plant Pathol 42:851–857 Manicom BQ, Bar-Joseph M, Kotze JM, Becker MM (1990) A restriction fragment length polymorphism probe relating vegetative compatibility groups and pathogenicity in Fusarium oxysporum f.sp. dianthi. Phytopathology 80:336–339 Manulis S, Kogan N, Reuven M, Ben-Yephet Y (1994) Use of the RAPD technique for identification of Fusarium oxysporum f.sp. dianthi from carnation. Phytopathology 84:98–101 McFadden HG, Wilson IW, Chapple RM, Dowd C (2006) Fusarium wilt (Fusarium oxysporum f.sp. vasinfectum) genes expressed during infection of cotton (Gossypium hirsutum). Mol Plant Pathol 7:87–101 Mes JJ, van Doom J, Roebroeck EJA, van Egmond E, van Aartrijk J, Boonekamp PM (1994) Restriction fragment length polymorphisms, races and vegetative compatibility groups within a worldwide collection of Fusarium oxysporum f.sp. gladioli. Plant Pathol 43:362–370 Mes JJ, Weststeijn EA, Herlaar F, Lambalk JJM, Wijbrandi J, Haring MA, Cornelissen BJC (1999) Biological and molecular characterization of Fusarium oxysporum f.sp. lycopersici divides race 1 isolates into separate virulence groups. Phytopathology 89:156–160 Miao VPW (1990) Using karyotype variability to investigate the origins and relatedness of isolates of Fusarium oxysporum f.sp. cubense. In: Ploetz RC (ed) Fusarium wilt of banana. The American Phytopathological Society, St Paul, MN, pp 55–62 Min XJ, Hickey DA (2007) Assessing the effect of varying sequence length on DNA barcoding of fungi. Mol Ecol Notes 7:365–373 Mishra PK, Fox RTV, Culham A (2003) Development of a PCR-based assay for rapid and reliable identification of pathogenic Fusaria. FEMS Microbiol Lett 218:329–332 Mouyna I, Renard JL, Brygoo Y (1996) DNA polymorphism among Fusarium oxysporum f.sp. elaeidis populations from oil palm, using a repeated and dispersed sequence “Palm”. Curr Genet 30:174–180 Mule G, Susca A, Stea G, Moretti A (2004) Specific detection of the toxigenic species Fusarium proliferatum and F. oxysporum from asparagus plants using primers based on calmodulin gene sequences. FEMS Microbiol Lett 230:235–240 Namiki F, Shiomi T, Kayamura T, Tsuge T (1994) Characterization of the formae specialis of Fusarium oxysporum causing wilts of cucurbits by DNA fingerprinting with nuclear repetitive DNA sequences. Appl Environ Microbiol 60:2684–2691 Neff M, Neff J, Chory J, Pepper A (1998) dCAPS, a simple technique for the genetic analysis of single nucleotide polymorphisms: experimental applications in Arabidopsis thaliana genetics. Plant J 14:387–392 Nelson KE (2003) The future of microbial genomics. Environ Microbiol 5:1223–1225 Nicholson P, Simpson DR, Weston G, Rezanoor HN, Lees AK, Parry DW, Joyce D (1998) Detection and quantification of Fusarium culmorum and Fusarium graminearum in cereals using PCR assays. Physiol Mol Plant Pathol 53:17–37 O’Donnell K, Kistler HC, Cigelnik E, Ploetz RC (1998) Multiple evolutionary origins of the fungus causing Panama disease of banana: concordant evidence from nuclear and mitochondria1 gene genealogies. Proc Natl Acad Sci USA 95:2044–2049 156 R. Saikia and N. Kadoo O’Donnell K, Kistler HC, Tacke BK, Casper HH (2000) Gene genealogies reveal global phylogeographic structure and reproductive isolation among lineages of Fusarium graminearum, the fungus causing wheat scab. Proc Natl Acad Sci USA 97:7905–7910 Paavanen-Huhtala S, Hyvonen J, Bulat SA, Yli-Mattila T (1999) RAPD-PCR, isozyme, rDNA RFLP and rDNA sequence analyses in identification of Finnish Fusarium oxysporum isolates. Mycol Res 103:625–634 Pasquali M, Marena L, Fiora E, Piatti P, Gullino ML, Garibaldi A (2004) Real-time PCR for the identification of a highly pathogenic group of Fusarium oxysporum f.sp. chrysanthemi on Argyranthemum frutescens L. J Plant Pathol 86:51–57 Pasquali M, Piatti P, Gullino ML, Garibaldi A (2006) Development of a real-time polymerase chain reaction for the detection of Fusarium oxysporum f.sp. basilici from basil seed and roots. J Phytopathol 154:632–636 Pasquali M, Dematheis F, Gullino ML, Garibaldi A (2007) Identification of race 1 of Fusarium oxysporum f.sp. lactucae on lettuce by inter-retrotransposon sequence-characterized amplified region technique. Phytopathology 97:987–996 Perez-Artes E, Roncero MIG, Diaz RMJ (1995) Restriction fragment length polymorphism analysis of the mitochondrial DNA of Fusarium oxysporum f.sp. ciceri. J Phytopathol 143:105–109 Ploetz RC (1990) Variability in Fusarium oxysporum f.sp. cubense. Can J Bot 68:1357–1363 Puhalla JE (1985) Classification of strains of Fusarium oxysporum on the basis of vegetative compatibility. Can J Bot 63:179–183 Purwantara A, Barrins JM, Cozijnsen AJ, Ades PK, Howlet BJ (2000) Genetic diversity of the Leptosphaeria maculans species complex from Australia, Europe and North America using amplified fragment length polymorphism analysis. Mycol Res 104:772–781 Schena L, Franco N, Ippolito A, Gallitelli D (2004) Real-time quantitative PCR: a new technology to detect and study phytopathogenic and antagonistic fungi. Eur J Plant Pathol 110:893–908 Schilling AG, Möller EM, Geiger HH (1996) Polymerase chain reaction-based assays for speciesspecific detection of Fusarium culmorum, F. graminearum, and F. avenaceum. Phytopathology 86:515–522 Schloss PD, Handelsman J (2005) Metagenomics for studying unculturable micro-organisms: cutting the Gordian knot. Genome Biol 6:229 Seifert KA, Samson RA, Dewaard JR, Houbraken J, Levesque CA, Moncalvo JM, Louis-Seize G, Hebert PD (2007) Prospects for fungus identification using CO1 DNA barcodes, with Penicillium as a test case. Proc Natl Acad Sci USA 104:3901–3906 Shimazu J, Yamauchi N, Hibi T, Satou M, Horiuchi S, Shirakawa T (2005) Development of sequence tagged site markers to identify races of Fusarium oxysporum f.sp. lactucae. J Gen Plant Pathol 71:183–189 Singh BP, Saikia R, Yadav M, Singh R, Chauhan VS, Arora DK (2006) Molecular characterization of Fusarium oxysporum f.sp. ciceri causing wilt of chickpea. Afr J Biotechnol 5:497–502 Sivaramakrishan S, Kannan S, Singh SD (2002) Genetic variability of fusarium wilt pathogen isolates of chickpea (Cicer arietinum L) assessed by molecular markers. Mycopathologia 155:171–178 Skovgaard K, Nirenberg HI, O’Donnell K, Rosendahl S (2001) Evolution of Fusarium oxysporum f.sp. vasinfectum races inferred from multigene genealogies. Phytopathology 91:1231–1237 Small J, Call D, Brockmn F, Straub T, Chandler D (2001) Direct detection of 16S rRNA in soil extracts by using oligonucleotide microarrays. Appl Environ Microbiol 67:4708–4716 Stewart JE, Kim MS, James RL, Dumroese RK, Klopfenstein NB (2006) Molecular characterization of Fusarium oxysporum and Fusarium commune isolates from a conifer nursery. Phytopathology 96:1124–1133 Tringe SG, von Mering C, Kobayashi A, Salamov AA, Chen KN, Chang HW et al (2005) Comparative metagenomics of microbial communities. Science 22:554–557 van der Does HC, Lievens B, Claes L, Houterman PM, Cornelissen BJC, Rep M (2008) The presence of a virulence locus discriminates Fusarium oxysporum isolates causing tomato wilt from other isolates. Environ Microbiol 10:1475–1485 7 Molecular Detection and Identification of Fusarium oxysporum 157 van der Nest MA, Steenkamp ET, Wingfield BD, Wingfield MJ (2000) Development of simple sequence repeat (SSR) markers in Eucalyptus from amplified inter-simple sequence repeats (ISSR). Plant Breed 119:433–436 van der Wolf JM, van Beckhoven JRCM, Bonanats PJM, Schoen CD (2001) New technologies for sensitive and specific routine detection of plant pathogenic bacteria. In: de Boer SH (ed) Plant pathogenic bacteria. Kluwer, Dordrecht, pp 75–77 Waalwijk C, De Koning JRA, Baayen RP, Gams W (1996) Discordant groupings of Fusarium spp from sections Elegans, Liseola and Dlaminia based on ribosomal ITS1 and ITS2 sequences. Mycologia 88:316–328 Wang PH, Lo1 HS, Yeh Y (2001) Identification of F. oxysporum cucumerinum and F. oxysporum luffae by RAPD-generated DNA probes. Lett Appl Microbiol 33:397–401 Wang B, Brubaker CL, Tate W, Woods MJ, Matheson BA, Burdon JJ (2006) Genetic variation and population structure of Fusarium oxysporum f.sp. vasinfectum in Australia. Plant Pathol 55:746–755 Ward E (1994) In: Blakemen JP, Williamson B (eds) Use of polymerase chain reaction for identifying plant pathogens. CAB International, Oxon, UK Whitehead DS, Coddington A, Lewis BG (1992) Classification of races by DNA polymorphism analysis and vegetative compatibility grouping in Fusarium oxysporum f.sp. pisi. Physiol Mol Plant Pathol 41:295–305 Woudt LP, Neuvel A, Sikkema A, van Grinsven MQJM, de Milliano WAJ, Campbell CL, Leslie JF (1995) Genetic variation in Fusarium oxysporum from cyclamen. Phytopathology 85:1348–1355 Wright KGF, Guest DI, Wimalajeewa DLS, Van Heeswijck R (1996) Characterisation of Fusarium oxysporum isolated from carnation based on pathogenicity, vegetative compatibility and random amplified polymorphic DNA (RAPD) assay. Eur J Plant Pathol 102:451–457 Zambounis AG, Paplomatas E, Tsaftaris AS (2007) Intergenic spacer–RFLP analysis and direct quantification of Australian Fusarium oxysporum f.sp vasinfectum isolates from soil and infected cotton tissues. Plant Dis 91:1564–1573 Zeller KA, Jurgenson JE, El-Assiuty EM, Leslie JF (2000) Isozyme and amplified fragment length polymorphisms (AFLPs) from Cephalosporium maydis in Egypt. Phytoparasitica 28:121–130 Zhang Z, Zhang J, Wang Y, Zheng X (2005) Molecular detection of Fusarium oxysporum f.sp. niveum and Mycosphaerella melonis in infected plant tissues and soil. FEMS Microbiol Lett 249:39–47 Chapter 8 Molecular Chemotyping of Fusarium graminearum, F. culmorum, and F. cerealis Isolates From Finland and Russia Tapani Yli-Mattila and Tatiana Gagkaeva Abstract PCR assays that yield markers that are predictive of NIV versus DON production or 3ADON versus 15ADON production were performed for 60 Fusarium graminearum isolates, most of which were from Finland and Russia. None of the F. graminearum isolates originating from Finland and north-western Russia produced any PCR fragment with DON and NIV specific primers, which indicates that they have a 3ADON chemotype. All F. graminearum isolates from southern Russia and most of the isolates from Asia and Germany produced fragments typical of the 15ADON chemotype. All F. culmorum isolates belonged to 3ADON chemotype, while all F. cerealis isolates belonged to NIV chemotype. The highest DON levels were found in the 3ADON molecular chemotype isolates of Finland, north-western Russia, and central Russia. In the combined 3ADON molecular chemotype isolates, DON production was clearly higher than in the combined 15DON chemotype isolates. 8.1 Introduction Fusarium head blight (FHB) caused by Fusarium graminearum (sexual state Gibberella zeae) and related Fusarium species is among the most important fungal disease of cereals worldwide (McMullen et al. 1997; Langseth et al. 1999; Bottalico and Perrone 2002; Gagkaeva and Yli-Mattila 2004; Ward et al. 2008). FHB T. Yli-Mattila Laboratory of Plant Physiology and Molecular Biology, Department of Biology, University of Turku, 20014 Turku, Finland e-mail: tymat@utu.fi T. Gagkaeva Laboratory of Mycology and Phytopathology, All-Russian Institute of Plant Protection (VIZR), 196608 St. Petersburg-Pushkin, Russia e-mail: t.gagkaeva@yahoo.com Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_8, # Springer-Verlag Berlin Heidelberg 2010 159 160 T. Yli-Mattila and T. Gagkaeva pathogens cause significant yield and quality losses and they pose a serious threat to food safety, because they produce mycotoxins. The most important mycotoxins are trichothecenes, such as deoxynivalenol (DON), nivalenol (NIV) and T-2, and zearalenone (ZEN). Multilocus molecular phylogenetic analyses have resolved the F. graminearum sensu lato as 13 phylogenetically distinct species (O’Donnell et al. 2000, 2004, 2008; Starkey et al. 2007; Yli-Mattila et al. 2009b). F. graminearum sensu stricto is the dominant FHB pathogen in most parts of Europe and North America (Láday et al. 2004; O’Donnell et al. 2000, 2004; Tóth et al. 2005; Yli-Mattila et al. 2009c). Recently F. graminearum has been spreading northward in Europe (e.g., Waalwijk et al. 2003; Nicholson et al. 2003) and has been replacing the closely related F. culmorum, which produces less DON than F. graminearum (Langseth et al. 1999; Jennings et al. 2004a; Jestoi et al. 2004, 2008). However, in some recent studies, F. culmorum was still recovered more frequently than F. graminearum in cereals e.g., in England, Wales, north-western Russia, and Finland based on percent contamination levels (Shipilova and Gagkaeva 1992; Jennings et al. 2004a, b). In Finland, FHB (Fig. 8.1) was first reported in the 1930s as a problem in oats (Rainio 1932). F. graminearum was reported in Finnish cereals as early as the 1960s (Ylimäki 1981), and subsequently in 1972 (Uoti and Ylimäki 1974); 1976– 1977 (Ylimäki et al. 1979); 1982–1984 (Rizzo 1993) and 1998 (Eskola et al. 2001; Yli-Mattila et al. 2002). The highest DON levels have been found in oat (Hietaniemi et al. 2008; Yli-Mattila et al. 2004a, b, 2008a, b, 2009a). High DON levels have also been found in spring wheat and barley, especially in central and western Finland, when harvesting has been delayed. The lowest levels have been found in winter rye and wheat, which are harvested early. Only low levels of ZEN Fig. 8.1 Symptoms on spring wheat (cultivar Mahti) 4 weeks after artificial inoculation in Finland 8 Molecular Chemotyping of Fusarium graminearum 161 have been found in Finnish field samples (Yli-Mattila et al. 2004a), although F. graminearum and F. culmorum are the main producers of ZEN (Bottalico and Perrone 2002; Jestoi et al. 2008). Based on quantitative TaqMan real-time PCR (qPCR) analysis of DNA recovered from cereal grains, F. graminearum is already more common than F. culmorum in Finland and a highly significant correlation has been found between F. graminearum DNA and DON in Finnish oat, barley, and spring wheat (Yli-Mattila et al. 2008a, b), which is in agreement with the results obtained in winter wheat in Sweden (Fredlund et al. 2008). In Finnish barley, F. culmorum also seems to contribute to DON production. The highly significant correlation between the level of F. graminearum/F. culmorum DNA and DON is in agreement with previous results of Waalwijk et al. (2004) and Nicholson et al. (2003) in Europe and Sarlin et al. (2006) in North America. In Russia, FHB caused by F. graminearum was first reported at the end of nineteenth century in the Far East region, where the environmental conditions are favorable to FHB (Voronin 1890). This fungus has not been detected in Siberia and the Ural regions, while in the southern Russia region (North Caucasus), where the first outbreak was reported in 1933 (Pronicheva 1935), F.graminearum is the most common causal agent of FHB. Since it was first detected in the North Caucasus region, FHB has reached epidemic levels several times in southern Russia during 1960s, 80s, and 90s and caused reduction of cereal yields (Ivanchenko 1960, Kirienkova 1992). In central Russia, where F. avenaceum tends to be the predominate cause of FHB, F. graminearum isolates have been found on the centralchernozem territory since 1980s (Selivanova et al. 1991). In the north-western region of Russia, this pathogen was not found until 2003, but since then this pathogen has become more common in this region (Gagkaeva et al. 2009). There are three distinct subpopulations of F. graminearum in the European part of Russia: southern, central, and north-western, which is adjacent to the Finnish population. The distance from the north-western and Finnish population to the southern population is ca. 2,000 km. Central population is located between them and is about 1,000 km from both of them. European F. graminearum populations are geographically separated from those in the Russian Far East and China by ca. 6,000 km (Fig. 8.2). F. graminearum isolates can be divided into two main groups (chemotypes) based on mycotoxin production and production profiles. One group produces DON and its acetylated derivatives and the other group produces NIV and its acetylated derivatives. The DON group can be further divided into 3ADON producers and 15ADON producers. The genetic basis of DON versus NIV production results from differences in Tri7 and Tri13 genes. Chemotype-specific PCR primers (Ward et al. 2002; Jennings et al. 2004a, b; Starkey et al. 2007) and multilocus genotyping that utilizes a suspension microarray (Ward et al. 2008) have been developed to predict the chemical phenotype (chemotype) of F. graminearum isolates. The chemotypes have also been called chemotype IA (producing 3ADON), IB (producing 15ADON) and II (producing NIV) (Miller et al. 1991). Trichothecene chemotype differences are apparently adaptive and trichothecene chemotype polymorphism has been 162 T. Yli-Mattila and T. Gagkaeva Fig. 8.2 The origin (black circles) of Fusarium isolates. The origin (black spots) of F. graminearum isolates in Far East and southern and central Russia have been marked with arrows and names. The three populations of F. graminearum in north-western Russia/Finland (A), southern and central Russia (B) and Far East (C) have been marked with black circles maintained through multiple speciation events by balancing selection (Ward et al. 2002). The 15ADON chemotype of F. graminearum is dominant in USA (Ward et al. 2008), England, and Wales (Jennings et al. 2004a, b) based on molecular and chemical analyses. Recent analyses of 3ADON populations in North America revealed that they had higher average growth rates, produced significantly more conidia, and accumulated significantly more trichothecene than isolates from sympatric 15ADON populations (Ward et al. 2008). NIV chemotype is relatively common e.g., in UK (Jennings et al. 2004b) and in some parts of southern USA (Gale et al. 2007). The chemical structures of DON and NIV differ from one another only by the presence or absence of a hydroxyl function at carbon atom 4 (C-4) of the core trichothecene molecule; NIV has the C-4 hydroxyl and DON does not. The trichothecene biosynthetic gene Tri13 is responsible for this structural difference; the Tri13 protein catalyzes the C-4 hydroxylation. In NIV-producing strains, Tri13 is functional and the C-4 position is hydroxylated, whereas in DON-producing strains the gene is non-functional due to multiple insertions and deletions in its coding region, and the C-4 position is not hydroxylated (Lee et al. 2002, Brown et al. 2002). Another trichothecene biosynthetic gene, Tri7, is responsible for acetylation of the C-4 hydroxyl and thereby is responsible for converting NIV to 4-acetylnivalenol (4-NIV). In DON-producing strains of F. graminearum, Tri7 is also nonfunctional due to multiple insertions and deletions (Lee et al. 2002; Brown et al. 2001). In at least some DON-producing strains, the nonfunctional Tri7 contains multiple repetitions of 11-bp sequence which occur only once in the functional Tri7 of NIV producers (Lee et al. 2001). 8 Molecular Chemotyping of Fusarium graminearum 163 In northern Europe, most F. graminearum and F. culmorum isolates previously studied possessed a 3ADON chemotype (Miller et al. 1991; Langseth et al. 1999; Chandler et al. 2003; Jestoi et al. 2004, 2008) based on chemical and molecular analysis, while in southern-European Russia the 15ADON chemotype was dominant (Leonov et al. 1990). In England and Sweden, the Tri7 gene of the trichothecene cluster was absent in all F. graminearum and F. culmorum isolates of the 3ADON chemotype (Chandler et al. 2003; Jennings et al. 2004a, b). In China also it was absent in all F. asiaticum (former F. graminearum lineage 6) isolates of the 3ADON chemotype (Zhang et al. 2007). A more specific way to separate 3ADON and 15ADON isolates is to use primers based on Tri3 gene (Ward et al. 2002, Jennings et al. 2004b). Genetic variation of Finnish and Russian F. graminearum isolates has been studied previously by different molecular methods (Yli-Mattila et al. 2004b, Gagkaeva and Yli-Mattila 2004) and by chemical analyses (Jestoi et al. 2004, 2008). In the present study, we reexamined most of these archived isolates as well as some new isolates, for two chemotype-specific PCR markers (Lee et al. 2001, Waalwijk et al. 2003), pathogenicity assays to wheat seedlings, and for DON and ZEN production in culture. The results were compared to those obtained with the previously described multilocus genotyping assay (Yli-Mattila et al. 2009a). 8.2 8.2.1 Materials and Methods Fusarium Isolates Isolates used in this study are listed in Tables 8.1 and 8.2 together with their geographic origin (Fig. 8.2), substrate, and results. Most isolates were obtained from small cereals grains. Grain samples were collected in different regions of Russia between 1998 and 2004, in Heilongjiang province (Harbin) of China in 1999 and in Germany in 1998. Six F. graminearum isolates from Finland were isolated from grain, root, and stem base. Six additional Finnish F. graminearum isolates of the years 2002 and 2003, used in molecular chemotyping, were obtained from grain (Table 8.1). Every isolate was single-spore subcultured and identified by morphological methods (Gerlach and Nirenberg 1982, Marasas et al. 1984, Figs. 8.3 and 8.4). The identification of each F. graminearum isolate was confirmed by PCR with species-specific primers Fg11f and Fg11r (Doohan et al. 1998, Waalwijk et al. 2003). At present all isolates (except for 6 Finnish F. graminearum isolates 02–1, 02–3, 02–11, 03–26, 03–27 and 03–28) are stored in the Collection of Laboratory of Mycology and Phytopathology, All-Russian Institute of Plant Protection (VIZR), Russia and ARS Culture Collection (NRRL), Peoria, IL. 164 T. Yli-Mattila and T. Gagkaeva Table 8.1 List of F. graminearum isolates used in mycotoxin (ng ml 1) and molecular chemotype (GzTri7f1/r1 and Tri13f/r primers) analyses. Six Finnish isolates of the years 2002 and 2003 are without NRRL number. Abbreviations: n.a. ¼ not analyzed, Ru ¼ Russia, CE ¼ central Russia, FE ¼ Far East, NW ¼ north-western Russia, SE ¼ southern Russia Origin (code for Finnish Number in NRRL isolates of the years 2002 and 2003) culture collection Host 45589 45590 Finland, Espoo Finland, Jalasjärvi 45595 45602 Finland, Pori Finland, Ylistaro 45845 45846 Finland, Isokyrö (02–06) Finland, Marttila (02-05) Finland, Isokyrö (02-11) Finland, Isokyrö (02-01) Finland, Isokyrö (02-03) Central Finland (03-26) Central Finland (03-27) Eastern Finland (03-28) Ru, CE, Bryansk Ru, CE, Bryansk Ru, CE, Bryansk Ru, CE, Tula Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Kamen-Ribolov Ru, FE, Khabarovsk Ru, FE, Khabarovsk Ru, FE, Khabarovsk Ru, FE, Ussuriysk Ru, FE, Ussuriysk Ru, NW, Leningrad Ru, NW, Leningrad Ru, NW, Leningrad Ru, NW, Leningrad Ru, NW, Leningrad Ru, NW, Leningrad Ru, NW, Leningrad Ru, SE, Krasnodar Ru, SE, Krasnodar Ru, SE, Krasnodar Ru, SE, Krasnodar Ru, SE, Krasnodar Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Barley Root Barley Stem base Wheat Root Oat Stem base Barley Grain Barley Grain Barley Grain Wheat Grain Wheat Grain Oat Grain Oat Grain Oat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Barley Grain Barley Grain Barley Grain Barley Grain Barley Grain Barley Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain Wheat Grain 45633 45634 45635 45636 45574 45615 45616 45617 45630 45575 45599 45605 45614 45579 45577 45606 45618 45628 45611 45612 45622 45623 45624 45625 45637 45584 45585 45586 45591 45600 45578 45580 45581 45582 45583 Tissue Year Toxins, mg ml 1 Chemotypespecific primers DON ZEN <4 1986 81 1986 138 251 GzTri7 Tri13 Suggested chemotype 3ADON 3ADON 1986 399 1993 104 120 11480 3ADON 3ADON 2002 2002 2002 2002 2002 2003 2003 2003 2004 2004 2004 2004 1998 2003 2003 2001 2003 1998 1998 1998 2003 1998 1998 1998 2003 2002 2003 2003 2003 2003 2004 2004 2004 1998 1998 1997 1997 1997 1998 1998 1998 1998 1998 n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. <4 28 188 <4 <4 107 27 25 741 795 237 <4 <4 251 <4 2510 43 2630 269 690 251 100 3550 245 1410 <4 <4 12880 <4 186 243 126 5010 126 3980 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON n.a. n.a. n.a. n.a. 3ADON 3ADON 3ADON 3ADON n.a. 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON n.a. 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON n.a. 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON 15ADON n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. 17 <4 29 258 15 50 13 12 142 <4 <4 16 20 50 8 45 80 50 80 224 102 317 200 564 170 20 <4 166 <4 8 36 32 63 <4 87 n.a. n.a. n.a. n.a. n.a. n.a n.a. n.a. n.a. þ þ þ þ þ þ þ þ n.a. n.a. þ þ þ þ n.a. þ þ þ þ þ þ þ þ þ þ þ þ n.a. n.a. þ þ þ þ þ þ þ (continued) 8 Molecular Chemotyping of Fusarium graminearum 165 Table 8.1 (continued) Origin (code for Finnish Number in NRRL isolates of the years 2002 and 2003) culture collection Host Tissue Year Toxins, mg ml 1 45596 45607 45608 45609 45610 45619 45620 45621 45626 45627 45629 45631 45632 45576 45587 45588 45593 45594 45603 45604 45592 45597 45613 45601 45598 Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia Ru, SE, North Ossetia China, Harbin China, Harbin China, Harbin China, Harbin China, Harbin China, Harbin China, Harbin Germany, Falkenhagen Germany, Falkenhagen Germany, Falkenhagen Germany, Reinshof Germany, Rocking Wheat wheat wheat wheat wheat Barley Barley Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain Grain 1998 2004 2004 2004 2004 2002 2002 2000 2004 2002 2002 2004 2004 1999 1999 1999 1999 1999 1998 1999 1998 1998 1998 1998 1998 8.2.2 DNA Extraction and PCR Chemotyping DON <4 <4 316 6 <4 8 24 21 10 <4 399 <4 32 12 <4 126 41 8 10 159 n.a. <4 n.a. <4 <4 ZEN 243 200 7080 1260 56 158 1200 200 64 76 <4 <4 23 562 1350 2510 1200 190 813 63 126 129 n.a. 295 447 Chemotypespecific primers Suggested chemotype GzTri7 Tri13 þ þ 15ADON þ þ 15ADON þ 15ADON þ þ 15ADON þ 15ADON þ þ 15ADON þ þ 15ADON þ þ 15ADON n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. þ þ 15ADON 3ADON þ 15ADON þ 15ADON þ þ 15ADON þ þ 15ADON 3ADON þ þ 15ADON þ þ NIV þ 15ADON þ þ 15ADON þ þ 15ADON The chloroform-octanol method was used for DNA extraction as described by Paavanen-Huhtala et al. (1999). Amplifications with chemotype-specific primers GzTri7f1/r1 (Lee et al. 2001) and Tri13f/r (Waalwijk et al. 2003) were performed for 60 F. graminearum isolates using a PTC-200 DNA Engine thermal cycler. According to previous investigations NIV producers generate a PCR fragment of 162 with primers GzTri7f1/r1 and a PCR fragment of 415 bp with primers Tri13f/r, while DON producers generate a PCR fragment of 162 bp plus a multiple of 11 bp with primers GzTri7f1/r1 and a PCR fragment of 234 bp with primers Tri13f/r. Amplifications were performed in 25-ml volumes containing Dynazyme reaction buffer (Finnzymes, Espoo, Finland), 150 mM each of dNTP and 1–10 ng of fungal DNA. The thermal cycler conditions used were as described by Lee et al. (2001), except that annealing was performed at 52 C. PCR reactions were repeated at least twice. PCR products from 23 isolates obtained with primers GzTrif1 and GzTrir1 were separated by electrophoresis in 2% MetaPhor agarose gel (FMC BioProducts, Rockland, ME, USA) in order to measure the length of the PCR product. 166 T. Yli-Mattila and T. Gagkaeva Table 8.2 F. culmorum, F. cerealis, and F. graminearum isolates and chemotypes identified by MLGT analyses. Abbreviations: Ru ¼ Russia, CE ¼ central Russia, FE ¼ Far East, NW ¼ north-western Russia, SE ¼ southern Russia NRRL # 45592 45597 45598 45601 45613 45642 45726 45727 45752 45758 45759 45765 45766 45770 45771 45774 45775 45776 45777 45778 45783 45784 45788 45803 45804 45829 45830 45831 45850 45851 45852 45854 45855 45856 45858 45859 45861 45897 45898 Origin Host Germany, Falkenhagen Germany, Falkenhagen Germany, Rocking Germany, Reinshof Germany, Falkenhagen Ru, FE, Khabarovsk Ru, NW, Arhangelsk Finland, Marttila Finland, Martilla Ru, CE, Moscow Ru, CE, Moscow China, Harbin China, Harbin China, Harbin Ru, CE, Moscow Ru, NW, Kaliningrad China, Harbin Ru, CE, Moscow Ru, Ural region, Bashkiria Ru, Ural, Bashkiria Ru, SE, Rostov Ru, SE, Rostov Ru, SE, North Ossetia Byelorussia Ru, NW, Pskov Kyrghyzstan Kyrghyzstan Ru, NW, Leningrad Finland, Marttila Finland, Marttila Finland, Marttila Finland, Marttila Western Finland, Etelä-Pohjanmaa Southern Finland, Uusimaa SW Finland, Satakunta SW Finland, Varsinais-Suomi Southern Finland, Uusimaa Western Finland, Isokyrö Finland, Marttila Wheat Wheat Wheat Wheat Wheat Wheat Solani Wheat Wheat Wheat Wheat Wheat Wheat Wheat Wheat Barley Wheat Wheat Wheat Wheat Cirsium sp. Cirsium sp. Cirsium sp. Wheat Cirsium sp. Cirsium sp. Cirsium sp. Cirsium sp. Barley Wheat Wheat Wheat Oat Tissue Year MLGT identification Grain 1998 graminearum Grain 1998 graminearum Grain 1998 graminearum Grain 1998 graminearum Grain 1998 graminearum Ear 2006 cerealis Potato 2002 culmorum Grain 2004 culmorum Grain 2004 culmorum Grain 2005 culmorum Grain 2005 culmorum Grain 2003 cerealis Grain 2003 cerealis Grain 2003 cerealis Grain 2004 culmorum Grain 2006 culmorum Grain 2003 cerealis Grain 2005 culmorum Root 2005 culmorum Root 2005 culmorum Leaf 2004 culmorum Leaf 2004 culmorum Leaf 2004 cerealis Ear 2003 culmorum Leaf 2004 culmorum Stem 2005 culmorum Stem 2005 culmorum Stem 2005 culmorum Grain 2002 culmorum Grain 2002 culmorum Grain 2002 culmorum Grain 2003 culmorum Grain 2003 culmorum Barley Oat Wheat Wheat Barley Barley Grain Grain Grain Grain Grain Grain 2003 2003 2003 2003 2001 2001 culmorum culmorum culmorum culmorum culmorum culmorum MLGT chemotype 15ADON NIV 15ADON 15ADON 15ADON NIV 3ADON 3ADON 3ADON 3ADON 3ADON NIV NIV NIV 3ADON 3ADON NIV 3ADON 3ADON 3ADON 3ADON 3ADON NIV 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 3ADON 8 Molecular Chemotyping of Fusarium graminearum 167 Table 8.3 Gene sequences amplified by multiplex PCR and ASPE (allele-specific primer extension, Ward et al. 2008) probes used for MLGT analysis Region Size (bp) Probe Target Reductase 702 RED-2 F. meridionale RED-ce F. cerealis RED-cu F. culmorum RED-p F. pseudograminearum RED-3 F. boothi RED-4 F. mesoamericanum RED-9 F. brasilicum Tri-101 911 AT-b B-FHB clade Fg complex AT-g F. austroamericanum AT-1 AT-2 F. meridionale AT-ce F. cerealis AT-cu F. culmorum AT-p F. pseudoamericanum AT-sp Fusarium sp. AT-3 F. boothi AT-4 F. mesoamericanum AT-5 F. acaciae-mearnsii AT-6 F. asiaticum AT-7 F. graminearum AT-8 F. cortaderiae EF 456 EF-g Fg complex F. mesoamericanum EF-1 EF-L F. lunulosporum EF-sp Fusarium sp. EF-5 F. acaciae-mearnsii EF-7 F. graminearum EF-8 F. cortaderiae MAT 1040 MAT-L F. lunulosporum MAT-6 F. asiaticum MAT-9 F. brasilicum Tri-3 912 T3-15ADON 15ADON T3-3ADON 3ADON T3-NIV NIV Tri-12 1163 T12-15ADON 15ADON T12-3ADON 3ADON T12-NIV NIV 8.2.3 Multilocus Genotyping The species and trichothecene chemotype composition of German F. graminearum isolates and F. culmorum and F. cerealis isolates were investigated using multiplex PCR with six primer pairs followed by a 37 probe version (Table 8.3) of six gene sequences of the multilocus genotyping (MLGT) assay. The products of the first PCR were used as templates in the multilocus genotyping assay. Multiplex amplifications, allele-specific primer extensions, microsphere hybridization, and 168 T. Yli-Mattila and T. Gagkaeva Fig. 8.3 Fusarium graminearum (left), F. cerealis (middle), and F. culmorum (right) conidia on SNA medium after 2 weeks in the dark Fig. 8.4 Fusarium graminearum (left), F. cerealis (middle) and F. culmorum (right) conidia on SNA medium after two weeks in the dark detection were performed as described by Ward et al. (2008). Hybridization and detection were performed using a Luminex 100 flow cytometer. 8.2.4 Mycotoxin Analyses A panel of 62 and 63 F. graminearum isolates was screened for DON and zearalenone (ZEN) production by indirect ELISA (enzyme linked immunosorbent assay, Kononenko and Burkin 2002). Every isolate was cultivated for 7 days in a small glass bottle (diameter 18 mm) with 1 ml PSA medium at 23 C. Then cultures of isolates were extracted with 1 ml of acetonitrile: water (6:1) and extracts were analyzed by ELISA, with detection limit 20 ng ml -1 of extract. 8.2.5 Pathogenicity Analysis The pathogenicity of F. culmorum (45726, 45776, 45777, 45784, 45803), F. graminearum (45618, 45633, 45702, 45710, 45720, 45744, 45827, 45636, 45638, 45713, 45762, 45799, 45832), and F. cerealis (45775, 45788, 45642) isolates was examined according to the modified method of Chelkowski and Manka (1983). 8 Molecular Chemotyping of Fusarium graminearum 169 Surface sterilized grains of winter wheat (cv. Moscowskay 39) were kept for one day in sterile water. Then healthy germinated grains were placed on the surface of the fungal colony, which was grown for one week on potato sucrose agar (PSA). The experiments were performed in 2 replicates, with 30 grains in three Petri dishes per replicate. In controls, 50 grains in five Petri dishes were placed on the surface of PSA medium in 2 replicates. The length of the seedlings and necrosis were estimated after one week of incubation in darkness at 23 C. The length of every seedling was measured and the mean of every isolate was compared to control. The symptoms of necrosis on the seedlings were evaluated using a scale with classes: 0 ¼ healthy seedlings, 1 ¼ small spots of necrosis on seedlings, 2 ¼ nearly 50% of the seedlings covered by brown lesions, 3 ¼ more than 50% of the seedlings have brown lesions or the seedlings are dead. 8.3 8.3.1 Results Molecular Chemotype Determination with Tri7 and Tri13 Primer Pairs When we analyzed the molecular chemotype results from the gel, we could find that none of the six F. graminearum isolates originating from north-western region of Russia and 12 isolates from Finland produced any PCR fragment with primers GzTri7f1/r1 and Tri13f/r (Tables 8.1 and 8.4). According to Kimura et al. (2003); Chandler et al. (2003) and Jennings et al. (2004a, b) the isolates which do not produce any PCR fragment with GzTri7f1/r primers belong to the chemotype 3ADON. So, it was possible to divide the isolates into 3ADON and 15ADON molecular chemotypes based on the presence or absence of amplicons, which in previous studies (Lee et al. 2001; Waalwijk et al. 2003) were shown to be markers for DON-producing isolates. Most isolates of the suggested 15ADON molecular chemotype yielded the DON-specific amplification fragment with GzTri7f1/r1 (31/34 isolates) and Table 8.4 Frequency of PCR fragments produced by F. graminearum isolates with chemotypespecific primers GzTri7f1/r1 and Tri13f/r Origin of isolates n % of isolates producing the PCR fragment, bp GzTri7f1/r1 Tri13f/r 161 (161 + x11) 412 234 Finland 12 0 0 0 0 North-western Russia 6 0 0 0 0 Southern Russia 18 0 89 0 83 Far East of Russia 12 0 75 0 50 China 7 0 71 0 71 Germany 5 20 80 20 60 Total 60 170 T. Yli-Mattila and T. Gagkaeva Tri13f/r primers (28/34 isolates). Six of the suggested 15ADON molecular chemotype isolates (NRRL 45579, 45577, 45585, 45581, 45582, and 45613) did not yield an amplicon with the GzTri7f1/r1 primer pair but did yield an amplicon with the Tri13f/r primer pair, while three isolates (NRRL 45608, 45610, and 45588) yielded an amplicon only with the primer pair Tri13f/r (Table 8.1). Most of the isolates from Asia (13/19 isolates), southern Russia (18/18 isolates), and Germany (4/5 isolates) produced fragments typical of the 15ADON molecular chemotype by at least one of the two primer pairs. Only one isolate (Germany, G.8-8) had an NIV molecular chemotype based on PCR products (Tables 8.1 and 8.4). The size of the PCR product obtained in 15ADON isolates with GzTri7f1/r1 primers was greatest in three isolates from North Ossetia collected in 1998 (206 bp). In the rest of 15ADON molecular chemotype isolates, the size was 173-195 bp (2–6 copies of the 11 bp repetition) and the biggest 15ADON group (11/22 isolates) had a PCR product of 184 bp. A single strain yielded an amplicon of 162 bp with primers GzTrif1/ri, and this is indicative of NIV production. The size variation of the PCR product between the 15ADON isolates was smaller than between the 50 isolates of Lee et al. (2001), who found 2–16 copies of the 11 bp repeat that occurs within Tri7. 8.3.2 Multilocus Genotyping The results of German F. graminearum isolates with MLGT assay were in accordance with the results obtained with molecular chemotype assays. All F. culmorum isolates possessed the 3ADON molecular chemotype. In contrast, all six isolates of F. cerealis possessed the NIV molecular chemotype (Table 8.2). The species identifications of the isolates in Table 8.2 could also be confirmed by using the MLGT assay. 8.3.3 DON and ZEN Production There were differences in mycotoxin production between F. graminearum isolates from different regions. Most F. graminearum isolates produced DON (47/62 isolates) and ZEN (51/63 isolates) (Table 8.1). The highest DON levels were produced by isolates with the 3ADON molecular chemotype that were from Finland and north-western Russia. Isolates from southern Russia with the 15ADON molecular chemotype produced lower levels of DON. Among all isolates with the 3ADON molecular chemotype, DON production was higher (154  39 ng ml 1) than among all isolates with the 15ADON molecular chemotype (37  11 ng ml 1). In contrast, there was no apparent difference in the levels of ZEN production between isolates with the 3ADON and 15ADON markers (1159  723 ng ml 1 8 Molecular Chemotyping of Fusarium graminearum 171 Table 8.5 DON and ZEN production (mean  SE) by F. graminearum isolates in 3ADON and 15ADON chemotypes Origin of isolates Chemotype (n) Toxin production, ng ml 1 DON ZEN Finland 3ADON (4) 180  74 2964  2839 851  546 North-western Russia 3ADON (6) 248  72 1831  808 Southern Russia 15ADON (18) 45  19 Far East of Russia 3ADON (4) 23  9 41  23 15ADON (8) 28  10 481  305 China 3ADON (2) 82 438 15ADON (5) 39  22 1055  399 254  93 <4 Germany 15ADON (2/3) <4 129 NIV (1) Total 50/51 Table 8.6 Pathogenicity of F. graminearum (15 ADON and 3 ADON chemotypes), F. culmorum (3ADON chemotype), and F. cerealis (NIV chemotype) to seedlings of winter wheat cv. Moscovskay 39 as compared to controls Necrosis, Fusarium sp. Chemotype No. of The length of seedlings as score compared to controls, % isolates F. graminearum 15 ADON 7 20.82  4.6 2.63  0.2 F. graminearum 3 ADON 6 11.93  2.6 2.94  0.05 F. culmorum 3 ADON 5 18.7  3.6 2.9  0.06 F. cerealis NIV 3 27.8  5.6 2.6  0.3 Control 100 0 in 3ADON isolates and 1362  438 ng ml 1 in 15ADON isolates), except in Russian Far East, where the isolates of the 15ADON chemotype produced clearly more ZEN than the combined isolates of 3ADON chemotype (Table 8.5), while in this area no clear difference was found in DON production between 3ADON and 15ADON isolates. The highest ZEN levels were produced by the F. graminearum isolates from southern Russia and by one Finnish isolate from oats (Table 8.1). 8.3.4 Pathogenicity of Isolates We analyzed the pathogenicity of isolates by determining their effect on seed germination, growth, and disease symptoms in the moderate resistant wheat cultivar Moscowskay 39. Seeds treated with all isolates of Fusarium resulted in significantly shorter seedlings than untreated, control seeds. In addition, large lesions were observed on seedlings resulting from treated seeds but not on seedlings resulting from untreated seeds (Table 8.6). F. graminearum isolates with 3ADON marker inhibited seed germination and reduced seedling growth significantly more than those with 15ADON marker. F. culmorum isolates with the 3ADON markers 172 T. Yli-Mattila and T. Gagkaeva also reduced seed germination and seedling growth more than F. graminearum isolates with 15ADON marker. The three F. cerealis isolates with NIV marker reduced seed germination and seedling growth less than the F. graminearum and F. culmorum isolates examined. The isolates examined exhibited a similar trend with respect to necrotic lesions. Isolates of F. graminearum and F. culmorum with the 3ADON molecular chemotype caused larger necrotic lesions on seedlings than isolates of F. graminearum with the 15ADON marker or isolates of F. cerealis with the NIV marker. 8.4 Discussion Although the 3ADON and 15ADON genetic markers of multilocus genotyping are highly correlated with trichothecene production profiles, there is no evidence that the genetic differences corresponding to the different PCR markers are the cause of the chemotypes with different production profiles. The Tri3 protein catalyzes acetylation of the C-15 hydroxyl, but Tri3 seems to be fully functional in both 3ADON and 15ADON producers. The Tri12 protein is an efflux pump that most likely pumps trichothecenes out of the cells in which they are synthesized (Desjardins 2006). So, it is not clear, how the genetic differences corresponding to the markers at Tri3 and Tri12 could lead to the differences in 3ADON and 15ADON production. This is in contrast to the Tri7 and Tri13 markers, for which it is quite clear that the genetic differences corresponding to the PCR markers directly cause the difference in the trichothecene production profiles. One F. graminearum strain from Far East has been shown to give a positive signal with the 3ADON probe from the Tri12 gene and with the 15ADON probe from the TRI3 gene suggesting that it may reflect recombination between isolates with these two chemotypes (Yli-Mattila et al. 2009b). The result is similar to the information of the article of Mirocha et al. (2003), according to whom Yoshizawa and Morooka (1973) had found a 3,15ADON isolate of F. graminearum in wheat and to the chromatographic analysis of one single-spore isolate from southern Russia, which produced both 3ADON and 15ADON based on chemical chromatographic analysis (Leonov et al. 1988). Kimura et al. (2003), Chandler et al. (2003) and Jennings et al. (2004a, b) have shown that isolates, which do not produce any PCR fragment with GzTri7f1/r primers belong to the chemotype 3ADON, because the Tri7 gene is deleted from all 3ADON isolates. Based on the results of the present work it also seems that isolates with the 3ADON molecular marker in most cases do not produce any PCR fragment with primers Tri13f/r. The only exception was one Chinese isolate (45603), which produced a PCR fragment with primers GzTri7f1/r and Tri13f/r. Ji et al. (2007) have also found Chinese 3ADON isolates of F. asiaticum and F. graminearum, which produced a PCR fragment with Tri13f/r primers. 8 Molecular Chemotyping of Fusarium graminearum 173 Forty-six of the 51 F. graminearum isolates tested with both the MLGT assay by Yli-Mattila et al. (2009b) and chemotype specific primer pair GzTri7f1/r1 in the present work produced concordant results. Ten isolates of Table 8.1 (45610, 45629, 45845, 45846, 02–01, 02–03, 02–11, 03–26, 03–27, 03–28) were not analyzed by MLGT assay. Most of the 15ADON molecular chemotype isolates produced the DON-specific amplification fragment both with GzTri7f1/r1 (31/34 isolates) and Tri13f/r primers (25/34 isolates). Seven of the 15ADON molecular marker chemotype isolates (isolates 45577, 45579, 45581, 45582, 45585, 45593, and 45613), which gave a positive result with GzTri7f1/r1 primers, did not produce the DON-specific amplification product with primers Tri13f/r. Only one F. graminearum isolate from China (isolate 45587), which gave a negative result with both primer pairs and was identified as a 3ADON molecular chemotype isolate in the present work, was later found to be a 15ADON molecular chemotype isolate based on MLGT assay (Yli-Mattila et al. 2009b). In addition, isolates 45588 from China and 45608 from North Ossetia having the 15ADON molecular marker (based on MLGT analysis) gave a negative result with primer pair GzTri7f1/r1. The 26 isolates of 3ADON chemotype (based on MLGT analysis) did not produce any amplification product, except for two isolates from North Ossetia (isolate 45596) and China (isolate 45603), which produced an amplification product both with Tri13f/r and GzTri7f/r primers. The molecular chemotype and MLGT results of the present work and those of Yli-Mattila et al. (2009b) are consistent with previous mycotoxin analyses of pure cultures of Finnish FHB isolates on autoclaved rice (Jestoi et al. 2004, 2008) and analyses of field samples (Yli-Mattila et al. 2008a, b), according to which Finnish F. graminearum and F. culmorum isolates belong to 3ADON chemotype. The idea that most 3ADON molecular chemotype isolates do not produce any PCR fragment with GzTri7f1/r and Tri13f/r primers is in agreement with the results of Waalwijk et al. (2003). According to Waalwijk et al. (2003) 26% of the Dutch F. culmorum isolates and 14% of the Dutch F. graminearum isolates did not produce any fragment with GzTri7f1/r primers, while 21% and 12% of the isolates of the same species did not produce any fragment with Tri13 primers. This probably means that ca. 20–25% of these Dutch F. culmorum and ca. 12–14% of the Dutch F. graminearum isolates belonged to the 3ADON molecular chemotype. Genotyping by chemotype-specific PCR may indicate only that mycotoxin gene is present, but it does not guarantee that the mycotoxin is produced or predict the levels of mycotoxin produced. According to the results of the present paper, 3ADON molecular genotypes produce usually more DON than 15ADON molecular genotypes, which is in agreement with the results obtained in certain areas of North America (Ward et al. 2008). According to Ward et al. (2008) and Yli-Mattila et al. (2008b, 2009b), 3ADON molecular chemotype frequencies among F. graminearum are increasing in certain areas of North America and Russian Far East, while in Russian Far East a new 3ADON producing species of F. graminearum species complex, Fusarium ussurianum, was recently found. In 1998 ten of 14 F. graminearum isolates collected in Russian Far East and Harbin in China were of the 15ADON molecular chemotype, while in 2006 twelve of 18 F. graminearum isolates had the 3ADON molecular chemotype (Yli-Mattila et al. 2009b). 174 T. Yli-Mattila and T. Gagkaeva 3ADON molecular chemotype isolates of F. graminearum and F. culmorum inhibited more strongly the growth of wheat seedlings and caused more necrotic lesions than 15ADON molecular chemotype isolates of F. graminearum and NIV isolates F. cerealis. This is in accordance with higher DON production of 3ADON molecular chemotype isolates in the present work and previous investigations (Ward et al. 2008) as compared to 15ADON molecular chemotype isolates. Only 3ADON molecular chemotype of F. graminearum was detected in the population of north-western Russia, where F. graminearum was not found until 2003. This population is probably related to the older 3ADON molecular chemotype population in Finland. Since, fitness is better (Ward et al. 2008) and 3ADON phytotoxicity and pathogenic potential are higher in 3ADON molecular chemotype isolates as compared to 15ADON molecular chemotype isolates, the increase of 3ADON chemotype of F. graminearum on new cereal production areas may be dangerous both for plant and mammal health. Acknowledgments We are indebted to Drs. A. Burkin and G. Kononenko from Laboratory of Mycotoxicology, All-Russian Institute for Veterinary Sanitation, Hygiene, and Ecology, Moscow, Russia for performing mycotoxins analyses. The visits of Dr. T. Yli-Mattila to the All-Russian Plant Protection Institute and the visits of Dr. T. Gagkaeva to the University of Turku were supported financially by the Academy of Finland (no. 126917 and 131957) and the Nordic network project New Emerging Mycotoxins and Secondary Metabolites in Toxigenic Fungi of Northern Europe (project 090014), which is funded by the Nordic Research Board. The authors are grateful to Drs. K. O’Donnell and T. Ward from USDA Research Institute in Peoria in USA for comments on the previous version of the manuscript and to Dr. Robert Proctor from the same institute for comments on the final version and for linguistic revision. References Bottalico A, Perrone G (2002) Toxigenic Fusarium species and mycotoxins associated with head blight in small-cereals in Europe. Eur J Plant Pathol 108:611–624 Brown DW, McCormick SP, Alexander HJ, Proctor RH, Desjardins AE (2001) A genetic and biochemical approach to study trichothecene diversity in Fusarium sporotrichioides and Fusarium graminearum. Fungal Genet Biol 32:121–133 Brown DW, McCormick SP, Alexander HJ, Proctor RH, Desjardins AE (2002) Inactivation of cytochrome P450 is a determinant of trichothecene diversity in Fusarium species. Fungal Genet Biol 36:224–233 Chandler EA, Simpson DR, Sthonsett MA, Nicholson P (2003) Development of PCR assays to Tri7 and Tri13 trichothecene biosynthetic genes, and characterisation of chemotypes of Fusarium graminearum and Fusarium cerealis. Phys Mol Plant Pathol 62:355–367 Chelkowski J, Manka M (1983) The ability of Fusaria pathogenic to wheat, barley and corn to produce zearalenone. Phytopathologische Zeitschrift-J Phytopathol 106:354–359 Desjardins AE (2006) Fusarium mycotoxins: Chemistry, Genetics and Biology. American Phytopathology Society, St Paul, MN Doohan FM, Parry DW, Jenkinson P, Nicholson P (1998) The use of species-specific PCR-based assays to analyse Fusarium ear blight of wheat. Plant Pathol 47:197–205 Eskola M, Parikka P, Rizzo A (2001) Trichothecenes, ochratoxin A and zearalenone contamination and Fusarium infection in Finnish cereal samples in 1998. Food Addit Contam 18:707–718 8 Molecular Chemotyping of Fusarium graminearum 175 Fredlund E, Gidlund A, Olsen M, Börjesson T, Spliid NHH, Simonsson M (2008) Method evaluation of Fusarium DNA extraction from mycelia and wheat for down-stream real-time PCR quantification and correlation to mycotoxin levels. J Microbiol Methods 73:33–40 Gagkaeva T, Yli-Mattila T (2004) Genetic diversity of populations of Fusarium graminearum in Europe and Asia. Eur J Plant Pathol 110:551–562 Gagkaeva TYu, Levitin MM, Nazarova LN, Sanin SS (2009) Infected grain and complex of Fusarium fungi on the territory of Russia Federation for 2004-2006 years. Agro XXI (n4–6): 4–6 Gale LR, Harrison SA, Milus EA, Ochocki JE, O’Donnell K, Ward TJ, Kistler, HC (2007) Diversity in Fusarium graminearum sensu stricto from the U.S: an update. In: Proceedings of the 2007 National Fusarium Head Blight Forum. The Westin Crown Center, Kansas City, Missouri 2-4 December 2007, p 26 Gerlach TR, Nirenberg H (1982) The genus Fusarium – a pictorial atlas. Kommissionsverlag Paul Parey, Berlin Hietaniemi V, Rämö S, Koivisto T, Pitkänen T, Ketoja E, Kartio M, Varimo K, Peltone S (2008) Viljojen mykotoksiinit Suomessa. In Happonen A (ed) Maataloustieteen Päivät 2008. Suomen Maataloustieteellisen Seuran tiedotteita no 23. Cited August 26, 2009. Published on January 9, 2008. ISBN 978-951-9041-51-3 (in Finnish) Ivanchenko MY (1960) Causal organisms of Fusarium head blight and their teleomorph stage in North Ossetia. Plant Protection 12:26–28 (in Russian) Jennings P, Coates ME, Turner JA, Chandler EA, Nicholson P (2004a) Determination of deoxynivalenol and nivalenol chemotypes of Fusarium culmorum isolates from England and Wales by PCR assay. Plant Pathol 53:182–190 Jennings P, Coates ME, Walsh K, Turner JA, Nicholson P (2004b) Determination of deoxynivalenol- and nivalenol-producing chemotypes of Fusarium graminearum isolates from wheat crops in England and Wales. Plant Pathol 53:643–652 Jestoi M, Paavanen-Huhtala S, Uhlig S, Rizzo A, Yli-Mattila T (2004) Mycotoxins and cytotoxicity of Finnish Fusarium strains grown on rice cultures. In: Canty SM, Boring T, Wardwell J, Ward RW (eds) Proceedings of the 2nd International Symposium on Fusarium Head Blight; incorporating the 8th European Fusarium seminar; 2004, 11–15; Orlando, Fl. USA. Michigan State University, East Lansing, MI, pp 405–409 Jestoi M, Paavanen-Huhtala S, Parikka P, Yli-Mattila T (2008) In vitro and in vivo mycotoxin production of Fusarium isolated from Finnish grains. Arch Phytopathol Plant Prot 41:545–558 Ji L, Cao T, Hu T, Wang S (2007) Determination of deoxynivalenol and nivalenol chemotypes of Fusarium graminearum isolates from China by PCR assay. J. Phytopathol 155:505–512 Kimura M, Tokai T, O’Donnell K, Ward TJ, Fujimura M, Hamamoto H, Shibata T, Yamaguchi I (2003) The trichothecene biosynthesis gene cluster of Fusarium graminearum F15 contains a limited number of essential pathway genes and expressed non-essential genes. FEBS Lett 539:105–110 Kirienkova AE (1992) Fusarium head blight of cereals in Krasnodar krai. In: Goncharov VT (ed) Fusarium head blight of cereals Krasnodar. 27:28 (in Russian) Kononenko GP, Burkin AA (2002) Fusarium toxins in grain food. Vete pathol 2:128–132 (in Russian) Láday M, Juhász Á, Mule G, Moretti A, Szécsi Á, Logrieco A (2004) Mitochondrial DNA diversity and lineage determination of European isolates of Fusarium graminearum (Gibberella zeae). Eur J Plant Pathol 110:545–550 Langseth W, Bernhoft A, Rundberget T, Kosiak B, Gareis M (1999) Mycotoxin production and cytotoxicity of Fusarium strains isolated from Norwegian cereals. Mycopathologia 144:103–113 Lee T, Han Y-K, Kim K-H, Yun S-H, Lee Y-W (2002) Tri13 and Tri7 determine deoxynivalenoland nivalenol-producing chemotypes of G. zeae. Appl Env Microbiol 68:2148–2154 Lee T, Oh DW, Kim H-S, Lee J, Kim Y-H, Youn S-H, Lee YW (2001) Identification of deoxynivalenol- and nivalenol-producing chemotypes of Gibberella zeae. Appl Environ Microbiol 67:2966–2972 176 T. Yli-Mattila and T. Gagkaeva Leonov AN, Kononenko GP, Soboleva NA (1990) Production of DON-related trichothecenes by Fusarium graminearum Schw from Krasnodarski krai of the USSR. Mycotox Res 6:54–60 Marasas WFO, Nelson PE, Toussoun TA (1984) Toxigenic Fusarium species: Identity and Mycotoxicology. Pennsylvania State University Press, University Park, Pennsylvania McMullen M, Jones R, Gallenberg D (1997) Scab of wheat and barley: a re-emerging disease of devastating impact. Plant Dis 81:1340–1346 Miller JD, Greenhalgh R, Wang Y, Lu M (1991) Trichothecene chemotypes of three Fusarium species. Mycologia 83:121–130 Mirocha CJ, Xie W, Filho ER (2003) Chemistry and detection of Fusarium mycotoxins. In: Leonard KJ, Bushnell WR (eds) Fusarium head blight of wheat and barley. APS Press, St. Paul, pp 144–164 Nicholson P, Chandler E, Draeger RC, Gosnan NE, Simpson DR, Thomsett M, Wilson AH (2003) Molecular Tools to Study Epidemiology and Toxicology of Fusarium Head Blight of Cereals. Eur J Plant Pathol 109:691–702 O’Donnell K, Kistler HC, Tacke BK, Casper HH (2000) Gene genealogies reveal global phylogeographic structure and reproductive isolation among lineages of Fusarium graminearum, the fungus causing wheat scab. Proc Natl Acad Sci USA 97:7905–7910 O’Donnell K, Ward TJ, Geiser DM, Kistler HC, Aoki T (2004) Genealogical concordance between the mating type locus and seven other nuclear genes supports formal recognition of nine phylogenetically distinct species within the Fusarium graminearum clade. Fungal Genet Biol 41:600–623 O’Donnell K, Ward TJ, Aberra D, Kistler HC, Aoki T, Orwig Kimura M, Bjørnstad Å, Klemsdal S (2008) Multilocus genotyping and molecular phylogenetics resolve a novel head blight pathogen within the Fusarium graminearum species complex from Ethiopia. Fungal Genet Biol 45:1514–1522 Paavanen-Huhtala S, Hyvönen J, Bulat SA, Yli-Mattila T (1999) RAPD-PCR, isozyme, rDNA RFLP and rDNA sequence analyses in identification of Finnish Fusarium oxysporum isolates. Mycol Res 103:625–634 Pronicheva LL (1935) Fusarium head blight of wheat in Azovo-Chernomorskom krai in 1934 and effect on yield. Plant Protection. 9:129:137 (in Russian) Rainio AJ (1932) Punahome Fusarium roseum Link. – Gibberella saubinettii (Mont.) Sacc. ja sen aiheuttamat myrkytykset kaurassa (Fusarium roseum beim Hafer und dadurch hervorgerufene Vergiftungen). Valtion Maatalouskoetoiminnan Julkaisuja 50:1–45 (in Finnish) Rizzo AF (1993) Determination of major naturally occurring Fusarium toxins in Finnish grains and feeds. The hemolytic activity of DON and T-2 toxin, and the lipid peroxidation induced in experimental animals. Ph.D. thesis, University of Helsinki Sarlin T, Yli-Mattila T, Jestoi M, Rizzo A, Paavanen-Huhtala S, Haikara A (2006) Real-time PCR for quantification of toxigenic Fusarium species in barley and malt. European J Plant Pathol 114:371–380 Selivanova TN, Baikova OV, Chernenko VYu (1991) Fusarium head blight in central-chernozem region. In Morosova LN (ed) Problems of protection of cereals against Fusarium head blight and others diseases. Minsk. 64:67 (in Russian) Shipilova NP, Gagkaeva TY (1992) Fusarium head blight in north-western region of Russia. Plant Prot. N11:7–8 (in Russian) Starkey DE, Ward TJ, Aoki T, Gale LR, Kistler HC, Geiser DM, Suga H, Toth B, Varga T, O’Donnell K (2007) Global molecular surveillance reveals novel Fusarium head blight species and trichothecene toxin diversity. Fungal Genet Biol 44:1191–1204 Tóth B, Mesterházy Á, Horváth Z, Bartók T, Varga M, Varga J. (2005) Genetic variability of central European isolates of the Fusarium graminearum species complex. Eur J Plant Pathol 113:35–45 Uoti J, Ylimäki A (1974) The occurrence of Fusarium species in cereal grain in Finland. Ann Agric Fenniae 13:5–17 8 Molecular Chemotyping of Fusarium graminearum 177 Voronin MC (1890) About tempulent corn in the South-Ussuriiskij region. Botanical notes. SPb, Russia. 3, 1:13–21 (in Russian) Waalwijk C, Kastelein P, de Vries I, van der Kerenyi Z, Lee T, Hesselink T, Kema G (2003) Major changes in Fusarium spp. in wheat in the Netherlands. Eur J of Plant Pathol 109:743–754 Waalwijk C, van der Heide R, de Vries I, van der Lee T, Schoen C, Costrel-de Corainville G, Häuser-Hahn I, Kastelein P, Köhl J, Lonnet P, Demarquet T, Kema GHJ (2004) Quantitative Detection of Fusarium Species in Wheat Using TaqMan. Eur J Plant Pathol 110:481–494 Ward TJ, Bielawski JP, Kistler HC, Sullivan E, O’Donnell K (2002) Ancestral polymorphism and adaptive evolution in the trichothecene mycotoxin gene cluster of phytopathogenic Fusarium. Proc Natl Acad Sci USA 99:9278–9283 Ward TJ, Clear R, Rooney A, O’Donnell K, Gaba D, Patrick S, Starkey D, Gilbert J, Geiser D, Nowicki T (2008) An adaptive evolutionary shift in Fusarium head blight pathogen populations is driving the rapid spread of more toxigenic Fusarium gramienarum in North America. Fungal Genet Biol 45:473–484 Ylimäki A (1981) The mycoflora of cereal seeds and some feedstuffs. Ann Agric Fenniae 20:74–88 Ylimäki A, Koponen H, Hintikka E-L, Nummi M, Niku-Paavola M-L, Ilus T, Enari TM (1979) Mycoflora and occurrence of Fusarium toxins in Finnish grain. Technical Research Centre of Finland, Materials and Processing Technology publication 21 (pp. 1–28). Valtion Painatuskeskus, Helsinki Yli-Mattila T, Paavanen-Huhtala S, Parikka P, Hietaniemi V, Jestoi M, Rizzo A (2004a) Toxigenic fungi and mycotoxins in Finnish cereals. In: Logrieco A, Visconti A (eds) An overview on toxigenic fungi and mycotoxins in Europe. Kluwer Academic Publishers, The Netherlands, pp 83–100 Yli-Mattila T, Paavanen-Huhtala S, Parikka P, Konstantinova P, Gagkaeva T (2004b) Molecular and morphological diversity of Fusarium species in Finland and northwestern Russia. Eur J Plant Pathol 110:573–585 Yli-Mattila T, Paavanen-Huhtala S, Parikka P, Konstantinova P, Gagkaeva T, Eskola M, Rizzo A (2002) Occurrence of Fusarium fungi and their toxins in Finnish cereals in 1998 and 2000. J Appl Genet 43A: 207–214 Yli-Mattila T, Paavanen-Huhtala S, Parikka P, Hietaniemi V, Jestoi M, Gagkaeva T, Sarlin T, Haikara A, Laaksonen S, Rizzo A (2008a) Real-time PCR detection and quantification of Fusarium poae, F. graminearum, F. sporotrichioides and F. langsethiae as compared to mycotoxin production in grains in Finland and Russia. – Arch. Phytopathol Plant Prot 41:243–260 Yli-Mattila T, O’Donnell K, Ward T, Gagkaeva T (2008b) Trichothecene chemotype composition of Fusarium graminearum and related species in Finland and Russia. J Plant Pathol 90:S60 Yli-Mattila T, Gagkaeva T, Ward TJ, Aoki T, Kistler HC, O’Donnell K (2009b) A novel Asian clade within the Fusarium graminearum species complex includes a newly discovered cereal head blight pathogen from the Far East of Russia. Mycologia (in press) Yli-Mattila T, Parikka P, Lahtinen T, Ramo S, Kokkonen M, Rizzo A, Jestoi M, Hietaniemi V (2009b) Fusarium DNA levels in Finnish cereal grains. In: Gherbawy Y, Mach L, Rai M (eds) Current Advances in Molecular Mycology. Nova Science Publishers, Inc., New York, pp 107–138 Yoshizawa T, Morooka N (1973) Deoxynivalenol and its monoacetate. New mycotoxins from Fusarium roseum and moldy barley. Agric Biol Chem 37:2933–2934 Zhang J-B, Li H-P, Dang F-J, Qu B, Xu Y-B, Zhao C-S, Lia Y-C (2007) Determination of the trichotchene mycotoxins chemotypes and associated geographical distribution and phylogenetic species of the Fusarium graminearum clade from China. Mycol Res 111:967–975 Chapter 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina: A Charcoal Rot Fungus Bandamaravuri Kishore Babu, Ratul Saikia, and Dilip K. Arora Abstract Macrophomina phaseolina is a global pathogen that inflicts losses on many agriculturally important crops worldwide, particularly in warm and tropical environments. Efforts to divide M. phaseolina into subspecies have been unsuccessful largely due to the extreme intraspecific variations in morphology and pathogenecity. The failure to adequately identify and detect M. phaseolina using conventional culture-based morphological techniques has led to the development of nucleic acid-based molecular approaches. PCR-based methods are highly sensitive and specific and have the potential to replace traditional technologies. Recently, species-specific oligonucleotide primers and digoxigenin (DIG)-labeled probe were designed at internal transcribed spacer (ITS) region for identification and detection of M. phaseolina. Accurate diagnosis and early detection of pathogens is an essential step in agriculture and environmental monitoring including plant disease management. The main objective of this review is to outline various molecular tools used for detection, identification, and characterization of M. phaseolina isolates. We also emphasize the significance of advanced technique such as real-time polymerase chain reaction (PCR) for qualitative and quantitative assays. B.K. Babu Environmental Microbiology Lab, Department of Environmental Engineering, Chosun University, Gwang ju-501759, South Korea e-mail: kishore_bandam@yahoo.co.in B.K. Babu and D.K. Arora National Bureau of Agriculturally Important Microorganisms (ICAR), Mau, Uttar Pradesh 275101, India e-mail: aroradilip@yahoo.co.in R. Saikia Biotechnology Division, North-East Institute of Science & Technology (CSIR), Jorhat 785006, Assam, India e-mail: rsaikia19@yahoo.com Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_9, # Springer-Verlag Berlin Heidelberg 2010 179 180 9.1 B.K. Babu et al. Introduction Macrophomina phaseolina (Tassi) Goid. is a primarily soilborne pathogen with wide distribution, varied host range, greater longevity, and higher competitive saprophytic ability (Chattanaver et al. 1988; Das 1988; Singh et al. 1988; Sobti and Bansal 1988; Abbaiah and Satayanarayan 1990; Das and Sankar 1990; Osunlaja 1990; Singh et al. 1990; Srivastava and Singh 1990; Kaur and Mukhopadhyay 1992; Siddiqui and Mahmood 1992; Mukherjee 1993). About 500 plant diseases are caused by the fungus (Su et al. 2001). The fungus is also associated with seeds and it has been shown that infection leads to both pre- and postemergence mortality, causing seed-to-seedling transmission of the pathogen (Pun et al. 1998). The pathogen occurs both inter- and intracellularly. The amount of internal inoculum is directly related with the degree of symptoms expressed (Sharma and Singh 2000). The pathogen attacks the root system of the plants and causes dissolution of all tissues except the xylem. Wilting followed by withering and death of aerial parts are characteristic symptoms of the disease (Edmunds 1964). The infected roots have abundant mycelia and sclerotia, but rarely pycnidia are produced on infected roots (Knox-Davies 1967). Both pycnidiospores and sclerotia have been implicated in the propagation of this fungus. The pathogen is plurivorus, causing ashy stem and blight or charcoal rot; root and stem show destruction of the cortex. The fungus is also the causal agent of the charcoal rot disease of many crops (Mihail 1992). In some cases, as in histological sections of roots infected with M. phaseolina showed destroyed cortical parenchyma, both giant cell and also cortical were invaded by mycelia (Suárez et al. 1998). Inhibition of penetration through the outer cell wall of the upper epidermis may be attributable to an osmophilic layer below the cell wall. Disruption of the host cell walls and subsequent host cell death were preceded by massive colonization of the host (Joye and Paul 1992). In this chapter, we will provide up-to-date information for the detection and identification of M. phaseolina. We will also discuss on molecular markers that have detected a wide range of genetic variations among the isolates. 9.2 Classification and Nomenclature Macrophomina phaseolina belongs to subdivision Duteromycotina, class Coelomycetes (Mihail 1992), order Sphaeropsidales, family Sphaerioidacease, and genus Macrophomina. The genus Macrophomina contains only one species: phaseolina (Sutton 1980). The successive changes in nomenclature led to confusion in adopting the correct name of Macrophomina. The monotypic genus Macrophomina was established by Petrak (1923) as M. philippinensis. Subsequently, Ashby (1927) examined the type material of this fungus, compared it with several other genera, and established earlier binomial for the fungus as Macrophomina phaseoli Maubl. Consequently, 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina 181 Ashby proposed the combination M. phaseoli (Maubl.) (Ashby) for M. phaseoli Maubl. and relegated the synonym M. cajani P. Syd. and Butler, M. chorchori Sawada, M. sesami Sewada, M. philippinensis Petrak, Sclerotium bataticola Taub., Rhizoctonia bataticola (Taub.) Butler, and Dothiorella cajani (P. Syd. and Butler) Petrak and H. Syd. Goidanich (1947) reviewed the taxonomy of Macrophomina. Petrak (1923) named this as M. phaseolina Tassi. in place of M. phaseoli Maubl. From 1947 onwards, the two names i.e., M. phaseoli (Maubl.) (Ashby) and M. phaseolina (Tassi.) Goid, became well established in psychopathological literature as the cause of charcoal rot of several important crop plants. After 1977, several other names were suggested for the fungus and ultimately Macrophomina phaseolina (Tassi.) Goid was accepted as the correct name (CMI description of pathogenic fungi and bacteria no. 275). The sclerotial phase of M. phaseolina is known as R. bataticola (Thakurji 1979; Punithalingam 1982). 9.3 Identification and Characterization The biggest problem before mycologists/plant pathologists is to identify/detect thousands of different isolates of this fungus from the cultures in infected roots, soil, and seed lots. Identification and detection of M. phaseolina is very difficult because the isolates are morphologically very similar. Different scientists have adopted different methods to distinguish M. phaseolina isolates. Among the methods that are most applicable are (1) morphological and cultural characterization, (2) biochemical methods, and (3) polymerase chain reaction (PCR) based molecular techniques. 9.3.1 Morphological and Cultural Characteristics M. phaseolina forms black colonies on potato dextrose agar (PDA) medium, and grows profusely at temperatures ranging from 15 to 40 C. However, the optimum growth occurs at 30–35 C. Some of the isolates can be identified on the basis of morphological characteristics and their thermophilic nature (Satto et al. 1999). The shape of the sclerotia, in most cases, is irregular except for a few which are round to oblong (Mandal et al. 1998). The mycelium is septate, 1.5–2.5 mm wide, hyaline at first turning to honey or black color. Fructification consists of globose or subglobose pycnidium, which is formed only on infected plants and consists of 3–4 layers of blackish-brown, thin-walled angular cells and sclerotia. Pycnidia can be detected in epidermis. Sclerotia are black, and their size on infected plants as well as on media is variable. Under laboratory conditions, sclerotium is hyaline to light brown in color measuring 89 mm in diameter, whereas in soil it measures from 50 to 120 mm in diameter (Upadhayay et al. 2002). 182 B.K. Babu et al. Insufficient morphological variability within the genus has led some workers to partition this fungus on the basis of cultural characteristics (Reichert and Hellinger 1947). Chromogenicity, sporulation ability, and pycnidial size are also known to diverge greatly (Crall 1948; Dhingra and Sinclair 1978; Pearson 1982). Traits with less variability would be more useful when trying to group the isolates. Different investigators have differentiated strains of this fungus on the basis of their ability to utilize nitrate as a nitrogen source (Correll et al. 1986; Correll et al. 1987; Larkin et al. 1988; Bayman and Cotty 1989). M. phaseolina have three types of growth patterns, viz., dense, feathery, and restricted, on chlorate-containing minimal medium (120 mM potassium chlorate). Earlier it was found that the utilization of chlorate was used as a marker for identifying host-specific isolates in M. phaseolina (Pearson et al. 1986; Cloud and Rupe 1991). Recent studies on the mechanism of chlorate assimilation and genetic basis for chlorate sensitivity have found no correlation with host specificity. It was also observed that chlorate-sensitive isolates were distinct from chlorate-resistant isolates within a given host (Su et al. 2001; Das et al. 2008). Therefore, the chlorate phenotype might not be useful for studying host specialization in M. phaseolina. 9.3.2 Biochemical and Serological Characterization Biochemical methods such as protein electrophoresis as well as fatty acid and isozyme profiles have also been applied for the characterization and differentiation of fungal taxa (Buth 1984; Faris et al. 1986). Immunological methods are highly specific and sensitive, which may provide a possible solution for detection and quantification of plant pathogenic fungi (Kitagawa et al. 1989; Eparvier and Alabouvette 1994; Jamaux and Spire 1994) and are reliable as quantitative techniques for ecological studies (Balesdent et al. 1995). However, there are no considerable reports on the use of biochemical as well as serological techniques to detect and quantify M. phaseolina. Srivastava and Arora (1997) have described a doubleantibody sandwich enzyme-linked immunosorbent assay (DAS-ELISA) technique using polyclonal antisera raised from soluble mycelial protein, cell wall proteins, and ribosomal proteins. Polyclonal antibodies that rose from soluble and cell wall protein found no significant polymorphism within the isolates of M. phaseolina. Ribosomal-specific antibodies detected in M. phaseolina at infected chickpea roots and use of these ribosomal-specific antibodies for the detection of M. phaseolina take place under some particular environmental conditions only. Therefore, this technique was limited to taxonomic identification and detection of M. phaseolina. 9.3.3 Molecular Methods for Characterization of M. phaseolina In recent years, there has been immense progress in the development of molecular biological tools and technologies. These have been increasingly applied to the study 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina 5.8S rRNA gene 18S IGS region 28S 183 5.8S rRNA gene ETS 28S 18S Structural genes Ribosomal RNA gene cluster ITSregions Primer ITS –1 5.8S rDNA 18S ribosomal RNA gene 28S ribosomal RNA gene Primer ITS – 4 ITS –1 region ITS – 2 region Fig. 9.1 General Physical map of rDNA gene cluster in fungal genome: The complete repeat unit is represented with genes location and spacer regions. ITS Internal transcribed spacers; IGS Intergenic spacers. primers ITS-1 and ITS-4 are showing the specific binding position of fungal plant pathogens (Leong and Holden 1989; Bridge and Arora 1998). The DNA sequences that encode for RNAs have been extensively used to study the taxonomic relationships and genetic variations in fungi (Bruns et al. 1992; Hibbert 1992). The ribosomal RNA gene cluster is found both in nucleus and mitochondria, and consists of both highly conserved and variable regions (White et al. 1990). The fungal nuclear rRNA genes are arranged as tandem repeats with several hundred copies per genome. The conserved sequences found in the large subunit and small subunit genes have been exploited to study the many relationships among distantly related fungi (Gaudet et al. 1989; Bowman et al. 1992; Bruns et al. 1992). The spacer regions between the subunits, called the internal transcribed spacers (ITS), and between the genes clusters, called the intergenic spacers (IGS), are considerably more variable than the subunit (Fig. 9.1). These genes have been used widely for studies on the relationships among species within a single genus or among interspecific populations (O’Donnel 1992; Molina et al. 1993; Li et al. 1994; Buscot et al. 1996; Arora et al. 1996; Singh et al. 2006). 9.3.3.1 Molecular Tools Used for Characterization of M. phaseolina Restriction Fragment Length Polymorphism Though there are many examples of the use of restriction fragment length polymorphism (RFLP) of spacer regions for discriminating between closely related species within a fungal genus, in the case of M. phaseolina the ITS region has shown no variations among isolates in restriction patterns of DNA fragments amplified by PCR covering the ITS region, 5.8S rRNA and part of 25S rRNA (Su et al. 2001; Babu et al. 2007). However, Purkayastha et al. (2006) found some degree of polymorphism in restriction patterns of the ITS region, including part of 25S 184 B.K. Babu et al. rDNA, indicating the RFLP analysis was not suitable for detection of genetic diversity of M. phaseolina. PCR-Fingerprinting Techniques Molecular fingerprinting techniques such as amplified fragment length polymorphism (AFLP), restriction fragment length polymorphism, random amplified polymorphic DNA (RAPD), and simple sequence repeats (SSR) have been used to unveil genetic variability in this soilborne filamentous fungi. Various studies have been devoted to the genetic diversity and pathogenic variability of M. phaseolina; however, only a single species has been recognized in the genus Macrophomina, but high levels of variation in pathogenecity have been observed (Mayek-Pe´rez et al. 2001; Su et al. 2001). Jana et al. (2003) developed taxonomic markers for population studies by using a single RAPD primer that distinguishes isolates of M. phaseolina from soybean, sesame, ground nut, chickpea, cotton, common bean, and 13 other hosts. The genetic diversity of M. phaseolina could favor its survival and adaptation to variable environments because of significant morphological (Mayek-Pe´rez et al. 1997), physiological (Mihail and Taylor 1995), pathogenic, and genetic (Chase et al. 1994; Vandemark et al. 2000; Mayek-Pe´rez et al. 2001; Pecina-Quintero et al. 2001; Su et al. 2001; Alvaro et al. 2003; Jana et al. 2003, 2005a, b) diversity. However, there is no clear evidence to suggest formae specialis, or subspecies. Recently, sorghum isolates of Indian origin were distinguished on the basis of chlorate sensitivity (Das et al. 2008). The correlation between genotype and geographical or biological origin was obtained with single RAPD marker among the Indian isolates of M. phaseolina collected from different hosts and various agroclimatic zones (Babu et al. 2009a). The Unweighted Pair Group Method with Arithmetic mean (UPGMA) dendrogram obtained by 10-mer random primer OPB08 separated 50 isolates physiological races has into 10 groups at 65–80% similarity (Fig. 9.2). The 10 clusters correlated well with the geographical locations, with exceptions for isolates obtained from Eastern Ghats (IV and X) and Western Ghats (VIII and IX) and Western Ghats (Fig. 9.2). There was a segregation of isolates from these two geographical locations into two clusters, thus distributing 10 genotypes into 8 geographical locations. The presence of two monomorphic bands suggests that the isolates might have evolved from a common ancestor but due to geographical isolation followed by natural selection and genetic drift they might have segregated into subpopulations (Babu et al. 2009a). Similarly, ReyesFranco et al. (2006) reported significant variations among the pathotypes obtained from different continents. 9.3.3.2 Diagnostic Tools Developed for Identification and Detection Molecular methods have recently been described to resolve genetic variation among the isolates of M. phaseolina (Das et al. 2008; Jana et al. 2005a, b; 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina mpK1 mpK2 mpK3 mpK8 mpK39 mpK5 mpK11 mpK18 mpK38 mpK40 mpK45 mpK32 mpK43 mpK9 mpK35 mpK36 mpK37 mpK6 mpK49 mpK50 mpK7 mpK14 mpK12 mpK13 mpK21 mpK23 mpK25 mpK19 mpK16 mpK44 mpK18 mpK42 mpK46 mpK47 mpK26 mpK27 mpK29 mpK30 mpK4 mpK17 mpK24 mpK41 mpK10 mpK20 mpK33 mpK15 mpK31 mpK34 mpK22 mpK28 0.6 0.7 0.8 0.9 I Western Himalyas II Gangetic plain III Semi Arid zone IV Eastern Ghats. a. V Central High lands Vl Decean Plateau Vlll Wastern Ghats .a. Vll North East lX Wastern Ghats .b. X Eastern Ghats .b. 185 1.0 Fig. 9.2 UPGMA-SAHN clustering dendrogram constructed by the data obtained from the primer OPB-8 in RAPD assay of Macrophomina phaseolina isolates labeled as 1–50 mpk. Geographical clusters I to X. Scale in the dendrogram shows the genetic similarity coefficient calculated according to Jaccard (1908).(Taken from Babu et al. 2009a) Purkayastha et al. 2006). These techniques are useful for grouping of isolates rather than their identification. Oligonucleotide-specific primers or probes targeting the ITS region have been reported to selectively detect several agriculturally important fungi such as Trichoderma, Hypocrea (Irina et al. 2005), Fusarium (Edel et al. 2000), Verticillium (Nazar et al. 1991), and Phytophthora (Lee et al. 1993). The most interesting thing about Macrophomina is that it has only one species (Sutton 186 B.K. Babu et al. 1980) but thousands of isolates. However, the screening of the GenBank for ITS sequences revealed the existence of very few sequences that showed some degree of variation among the isolates. Sequence variation in the rRNA genes may allow the use of these genes as targets for differential amplification. PCR-Based Identification ITS-RFLP of M. phaseolina could not detect variations within the different isolates (Su et al. 2001, Babu et al. 2007) (Fig. 9.3). Sequencing and alignment of eight isolates collected from different hosts and diverse ecological conditions, along with two sequences from the GenBank, revealed various conserved and variable regions in the ITS sequences. For a better understanding of these regions, the amplified ITS sequence was virtually divided into five regions (Fig. 9.4). Region 4 (not depicted in M 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 M Fig. 9.3 RFLP analysis of ITS region digested with HpaI showing similar restriction pattern with different Macrophomina phaseolina isolates (Taken from Babu 2008) Internal transcribed spacer 1 Internal transcribed spacer 2 AF 132795 : U 97333 : DQ359737 : DQ359738 : DQ359739 : DQ359740 : DQ359741 : DQ359742 : DQ359743 : DQ359744 : Region 1 110 bp Region 2 Region 3 330 bp Region 5 402 bp Fig. 9.4 Development of specific oligonucleotide primers and probe: Alignment of ITS-1 and ITS-2 sequences from eight isolates of Macrophomina phaseolina and two reference sequences (AF132795 and U97333) from GenBank database. Nucleotides are shown as color bars (A-red, G-yellow, T-blue and C-green). Regions 1, 2, 3 and 5 are completely aligned, 5.8S RNA gene is not shown and the region 4 omitted because of the variability. Solid line rectangles indicate specific nucleotide areas used for the development of specific oligonucleotide primers and the dashed ones show the specific region used for probe designing. The position of the first nucleotide of region 1 and others are according to the reference sequence AF132795 (Taken from Babu et al. 2007) 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina 187 Table 9.1 Species-specific oligonucleotides Sl. No. Primer/Probe Sequence 1 MPKF1 (primer) 50 -CCG CCA GAG GAC TAT CAA AC-30 2 MPKR1(primer) 50 -CGT CCG AAG CGA GGT GTA TT-30 3 MPKH1 (probe) 50 -GCT CTG CTT GGT ATT GGG C-30 the figure) was deleted, as it contained sequences that showed variations among the isolates of M. phaseolina. Further sequence alignment with other closely related genera of soilborne fungi helped to identify two regions that were conserved among Macrophomina isolates but exhibited a high degree of variability among isolates of other genera. The two primers designed from these selected regions of ITS showed specificity in the PCR assay. Optimization of the PCR conditions and validation of primers yielded a specific 350-bp amplicon for M. phaseolina isolates. The absence of the 350-bp product in other species of soilborne fungi, bacteria, and actinomycetes confirmed that the primers can be utilized to selectively and specifically identify M. phaseolina. Thus PCR assays with primers MpKR1 and MpKF1 could be used to rapidly identify M. phaseolina. Hybridization Probes Further, species-specific oligonucleotide probe MpKH1 was also designed within the ITS region (Table 9.1) (Babu et al. 2007). The DIG-labeled probe could detect the target sequences at varying concentrations with little or no background. The probe was also shown to be specific for M. phaseolina and no signals were obtained with nonspecific target ITS sequences (Fig. 9.4). Though the probe would be a good diagnostic tool for the detection within the PCR-amplified product of pure cultures, in direct field detection with soil DNA it had shown false-positive signal and nonspecific binding with closely related species (Babu 2008). Therefore, hybridization assays developed for M. phaseolina based on ITS region have been limited to certain specific conditions. 9.4 Recent Developments in the Diagnostics of M. phaseolina Previously, we demonstrated molecular identification and detection tools for M. phaseolina (Babu et al. 2007) in which rDNA gene cluster had been selected as a target for designing of specific primers and probe. Even though primers were showing greater specificity, hybridization assay using DNA probe was not always sensitive because of its small size (20 bp). Furthermore, like post-PCR analysis, dot-blot techniques are very time-consuming and require additional skills. The diagnostic technique has greatly improved by the introduction of real-time PCR technology based on fluorescence detection and quantification during PCR 188 B.K. Babu et al. amplification. Modern quantitative (qPCR) technology employing either nonspecific fluorogenic DNA-binding dyes, such as SYBR Green, or sequence-specific fluorogenic hybridization probes, such as TaqMan (Cook and Schlitzer 1981; Guiver et al. 2001) detection chemistry, has been described for a range of plant pathogens (Falcao et al. 1993; Widjojoatmodjo et al. 1999; Arvanitidou et al. 2000; Hao-Zhi and Ruey 2006) including assays for the detection both in the laboratory (Doggett 2000; Stultz et al. 2001) and in the field (Spencer and Spencer 1997). Therefore, recently we developed real-time qPCR assay to quantify M. phaseolina abundance in soils and plants by using TaqMan and SYBR Green assay techniques. The techniques were targeting on ~1 kb sequence characterized amplified region (SCAR) of M. phaseolina, and two sets of specific primers were designed for SYBR green (MpSyK) and TaqMan (MpTqK) assays. Both the assays were reproducible and accurately quantified M. phaseolina population in soils and plant samples. No cross hybridization and no increasing fluorescent signal exceeding the baseline threshold were observed with the tested microbes, except when M. phaseolina DNA was used as template. Limit of quantification (LOQ) of M. phaseolina DNA in sclerotial suspension was calculated as 200 pg/ml equivalent to 1 ng, which is equivalent to 2  104 CFU g–1 per soil. Further, we demonstrated the application of a species-specific real-time qPCR assay useful for the detection of M. phaseolina population in soil (Babu et al. 2009b). 9.5 Future Prospects Sequence data in public databases are constantly increasing; as a result, integration of more strains into detection systems of M. phaseolina will become possible, and identification of this pathogen is likely to become an easier task which would also help current detection and characterization methods. The resulting database will allow the complete analysis of developmental processes that are characteristics of the fungus, including the molecular nature of pathogenecity. Like some other phytopathogenic fungi, in the next few years complete genome sequences might be available for M. phaseolina also, and a combination of DNA microarrays with other genomic methods will certainly accelerate the effort to characterize this fungus. The escalating effect of biology, bioinformatics, and genomics may facilitate tracing out M. phaseolina genetic resources and their potential application in disease management also. 9.6 Conclusion In this review, an overview of our current knowledge regarding the biology of M. phaseolina with emphasis on diagnosis is given. Although several attempts have been made to examine the pathogenic and genetic variability of the species, further 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina 189 work is needed to clarify, for example, the adaptability to multiple hosts and the validity of species evolution in the genus Macrophomina. Similarly, our knowledge on the PCR identification and hybridization assays for detection has been enhanced. Therefore, further studies should be pursued to quantify the population of this important plant pathogen in the field. Besides, diagnosis genes involved in the pathogenesis and characterization of host-specific toxins would hasten efficient breeding and the development of resistance traits. Acknowledgment The work is supported by a Network Project, sponsored by Indian Council of Agricultural Research (ICAR), Government of India, New Delhi. The authors are thankful to the funding agency. References Abbaiah K, Satayanarayan A (1990) Stem canker on black gram (Vigna mungo) in Andhra Pradesh. Indian J Plant Protect 18:136 Alvaro MRA, Ricardo VA, Carlos AAA, Valdemar PC, David SJF, Silvana RRM, Luis CB, Mauro CP, Claudio GPC (2003) Genotypic diversity among Brazilian isolates of Macrophomina phaseolina revealed by RAPD. Fitopatol Brasil 28:279–285 Arora DK, Hirsch PR, Kerry BR (1996) PCR-based molecular discrimination of Verticillium chlamydosporum isolates. Mycol Res 100:801–809 Arvanitidou M, Spaia S, Velegraki A, Pazarloglou M, Kanetidis D, Pangidis P, Askepidis N, Katsinas C, Vayonas G, Katsouyannopoulos V (2000) High level of recovery of fungi from water and dialysate in haemodialysis units. J Hosp Infect 45:225–230 Ashby SF (1927) Macrophomina phaseolina (Maubl.) Comb. Nov. The pycnidial stage of Rhizoctonia bataticola (Taub.). Butl Trans Br Mycol Soc 12:141–147 Babu KB (2008) Genomic fingerprinting and development of molecular probes for identification and detection of Macrophomina phaseolina. Ph.D. Thesis, Bundelkhand University, India Babu KB, Srivastava AK, Saxena AK, Arora DK (2007) Identification and detection of Macrophomina phaseolina by using species-specific oligonucleotide primers and probe. Mycologia 99:733–739 Babu KB, Reddy SS, Sukumar M, Arora DK (2009b) SCAR derived, Quantitative Real-Time PCR assay for enumeration of phytopathogenic fungi - Macrophomina phaseolina. Appl Environ Microbiol. unpublished data Babu KB, Reddy SS, Yadav KM, Sukumar M, Mishra V, Saxena AK, Arora DK (2009a) Geodiversity among Indian isolates of Macrophomina phaseolina by using RAPD marker. Indian J Microbiol (In press) Balesdent MH, Destheux C, Gall PR, Rouxel T (1995) Quantification of Leptosphaeria maculans growth in cotyledons of Brassica napus using ELISA. J Phytopathol 143:65–73 Bayman P, Cotty PJ (1989) Vegetative compatibility groups in Aspergillus flavus. Phytopathology 79:1186 (Abstr.) Bowman BH, Taylor JW, Browilee AG, Lee J, Lu S, White TJ (1992) Molecular evolution of the fungi: relationship of basidiomyetes and chytridiomycetes. Mol Biol Evol 9:285–296 Bridge PD, Arora DK (1998) Interpretation of PCR methods for species definition. Application of PCR in mycology. CAB International, pp 63–84 Bruns TD, Vilgalys R, Barns SM, Gonzalez D, Hibbert DS, Lane DJ, Simon LS, Szaro TM, Weisburg WG, Sogin ML (1992) Evolutionary relationships within fungi: analysis of nuclear small subunit rRNA sequences. Mol Phyl Evol 1:231–241 190 B.K. Babu et al. Buscot F, Wipf D, Di Battista C, Munch JC, Botton B, Martin F (1996) DNA polymorphism in morels: PCR-RELP analysis of the ribosomal DNA spacers and microsatellite-primed PCR. Mycol Res 100:63–71 Buth DG (1984) The application of electrophoretic data in systematic studies. Annu Rev Ecol Syst 15:501–522 Chase TE, Jiang Y, Mihail JD (1994) Molecular variability in Macrophomina phaseolina. Phytopathology 84:1149 Chattanaver SN, Adiver SS, Nargound VB, Rao SS, Kulkarni S (1988) Laboratory evaluation of fungicide against Rhizoctonia bataticola causing root rot of Casuarina. Curr Res 17:160–161 Cloud GL, Rupe JC (1991) Morphological instability in a chlorate medium of isolates of Macrophomina phaseolina from soybean and sorghum. Phytopathology 81:892–895 Cook WL, Schlitzer RL (1981) Isolation of Candida albicans from freshwater and sewage. Appl Environ Microbiol 41:840–842 Correll JC, Puhalla JE, Schneider RW (1986) Identification of Fusarium oxysporum f.sp.apii on the basis of colony size, virulence, and vegetative compatibility. Phytopathology 76:396–400 Correll JC, Klittich CJR, Leslie JF (1987) Nitrate nonutilizing mutants of Fusarium oxysporum and their use in vegetative compatibility tests. Phytopathology 77:1640–1666 Crall JM (1948) Charcoal rot of soybean caused by Macrophomina phaseoli (Maubl.) Ashby. Ph. D. dissertation, University of Missouri, Columbia, p 148 Das ND (1988) Effect of different sources of carbon, nitrogen and temperature on growth and sclerotial production of Macrophomina phaseolina (Tassi) Goid, causing root rot/charcoal rot disease of castor. Indian J Plant Pathol 6:97–98 Das ND, Sankar GRM (1990) Effect of root rot disease caused by Macrophomina phaseolina Tassi (Goid). Of castor (Ricinus comunis) by statistical modeling of cultural practices of carbendazim applications. Indian J Mycol Plant Pathol 20:234–240 Das IK, Fakrudin B, Arora DK (2008) RAPD cluster analysis and chlorate sensitivity of some Indian isolates of Macrophomina phaseolina from sorghum and their relationships with pathogenicity. Microbiol Res 163:215–224 Dhingra OD, Sinclair JB (1978) Biology and pathology of Macrophomina phaseolina Vicosa. Imprensia Universitaria, Universidade Federal de Vicosa, Brazil, pp 166 Doggett MS (2000) Characterization of fungal biofilms within a municipal water distribution system. Appl Environ Microbiol 66:1249–1251 Edel V, Steinberg C, Gautheron N, Alabouvette C (2000) Ribosomal DNA-targeted oligonucleotide probe and PCR assay specific for Fusarium oxysporum. Mycol Res 104:518–526 Edmunds LK (1964) Combined relation of plant maturity, temperature, and soil moisture to charcoal stalk rot development in grain sorghum. Phytopathology 54:514–517 Eparvier A, Alabouvette C (1994) Use of ELISA and GUS-transformed strains to study competition between pathogenic and non-pathogenic Fusarium oxysporum for root colonization. Biocont Sci Technol 4:35–47 Falcao DP, Leite CQF, Simoes MJS, Giannini MJSM, Valentini SR (1993) Microbial quality of recreational waters in Araraquara, SP, Brazil. Sci Total Environ 128:37–49 Faris MA, Sabo FE, Cloutier Y (1986) Intraspecific variation in gel electrophoresis patterns of soluble mycelial proteins of Phytophthora megasperma isolated from alfa alfa. Can J Bot 64:262–265 Gaudet J, Julien J, Lafay JF, Brygoo Y (1989) Phylogeny of some Fusarium species, as determined by large subunits rRNA sequence comparison. Mol Biol Evol 6:227–242 Goidanich (1947) Differential responses of Guayule (Parthinium arginatum gray) genotypes to sulture fintrate and toxin from Macrophomina phaseolina (Tassi) Goid. Kuti. Jo; Schading RL, Latigo GC, Bradfold JM. Phytopathologisete Zeitschrift 145(7):305–311 Guiver M, Levi K, Oppenheim BA (2001) Rapid identification of Candida species by TaqMan PCR. J Clin Pathol 54:362–366 Hao-Zhi Y, Ruey FL (2006) Selection of internal control genes for real-time quantitative RTPCR assays in the oomycete plant pathogen Phytophthora parasitica. Fungal Genet Biol 43:430–438 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina 191 Hibbert DS (1992) Ribosomal RNA and fungal systematics. Trans Mycol Soc Jpn 33:533–556 Irina SD, Alexei GK, Monika K, John B, George S, Christian PK (2005) An oligonucleotide barcode for species identification in Trichoderma and Hypocrea. Fungal Genet Biol 42:813–828 Jaccard P (1908) Nouvelles researches sur la distribution florale. Société Vaudoise des Sci Naturelles Bull 44:223–70 Jamaux I, Spire D (1994) Development of polyclonal antibody based immunoassay for the early detection of Selerotina selerotioum in rapeseed petals. Plant Pathol 43:847–862 Jana TK, Sharma TR, Prasad RD, Arora DK (2003) Molecular characterization of Macrophomina phaseolina and Fusarium species by a single primer RAPD technique. Microbiol Res 158:249–257 Jana TK, Sharma TR, Singh NK (2005a) SSR-based detection of genetic variability in the charcoal root rot pathogen Macrophomina phaseolina. Mycol Res 109:81–86 Jana TK, Singh NK, Koundal KR, Sharma TR (2005b) Genetic differentiation of charcoal rot pathogen, Macrophomina phaseolina, in to specific groups using URP-PCR. Can J Microbiol 51:159–164 Joye GF, Paul RN (1992) Histology of infection of Hydrilla verticillata by Macrophomina phaseolina. Weed Sci 40:288–295 Kaur NP, Mukhopadhyay AN (1992) Integrated control of chickpea wilt complex by Trichoderma and chemical methods in India. Trop Pest Manag 38:372–375 Kitagawa T, Sakamoto Y, Furumi K, Ogura H (1989) novel enzyme immunoassay for specific detection of Fusarium oxysporum F.sp. cucumerinum and for general detection of various Fusarium species. Phytopathology 79:162–165 Knox-Davies PS (1967) Mitosis and aneuploidy in the vegetative hyphae of Macrophomina phaseolina. Am J Bot 54:1290–1295 Larkin RP, Hopkins DL, Martin FN (1988) Differentiation of strains and pathogenic races of Fusarium oxysporum F.sp. niveum based on vegetative compatibility. Phytopathology 78:1542 (Abstr.) Lee SB, White TJ, Taylor JW (1993) Detection of Phytophthora species of oligonucleotide hybridization to amplified ribosomal DNA spacers. Phytopathology 83:177–181 Leong SA, Holden DW (1989) Molecular genetic approaches to the study of fungal pathogenesis. Annu Rev Phytopathol 27:463–481 Li KN, Rouse DI, German TL (1994) PCR primers that allow intergeneric differentiation oa ascomycetes and their application to Verticillium spp. Appl Environ Microbiol 60:4324–4331 Mandal R, Sinha MK, Ray MKG, Mishra CBP, Chakrabarty NK (1998) Variation in Macrophomina phaseolina (Tassi) Goid causing stem rot of jute. Environ Ecol 16:424–426 Mayek-Pérez N, López-Castañeda C, Acosta-Gallegos JA (1997) Variación en caracterı́sticas culturales in vitro de aislamientos de Macrophomina phaseolina y su virulencia en frijol. Agrociencia 31:187–195 Mayek-Pérez N, Lopez-caataneda C, Gonzalez-Chavira M, Garch-Espinosa R, Acosta-Gallegos J, de la Vega OM, Simpson J (2001) Variability of Mexican isolates of Macrophomina phaseolina based on pathogenesis and AFLP genotype. Physiol Mol Plant Pathol 59:257–264 Mihail JD (1992) In Methods for research on soilborne phytopathogenic fungi. In: Singleton LL, Mihail JD, Rush CM (eds) St Paul, MN, USA, American Phytopathological Society Press, pp 134–136 Mihail JD, Taylor SJ (1995) Interpreting variability among isolates of Macrophomina phaseolina in pathogenicity, pycnidium production and chlorate utilization. Can J Bot 10:1596–1603 Molina FI, Jong S, Huffman JL (1993) PCR amplification of the 3 external transcribed and intergenic spacers of the ribosomal DNA repeat unit in three speices of Saccharomyces. FEMS Microbiol Lett 108:259–264 Mukherjee PK (1993) Biological control of collar rot, dry root rot wilt and gray mold of chickpea. Patencheru AP, India, p 36 192 B.K. Babu et al. Nazar RN, Hu X, Schmidt J, Culham D, Robb J (1991) Potential use of PCR-amplified ribosomal intergenic sequences in the detection and differentiation of Verticillium wilt pathogens. Physiol Mol Plant Pathol 39:1–11 O’Donnel KL (1992) Ribosomal DNA internal transcribed spacers are highly divergent in the phytopathogenic ascomycetes Fusarium sambucinum (Gibberella pulicaris). Curr Genet 22:213–220 Osunlaja SD (1990) Tillage effects on incidence and severity of root and stalk rot caused by Macrophomina phaseolina in South West Nigeria. J Phytopathol 130:312–316 Pearson CAS (1982) Macrophomina phaseolina Casual organism of charcoal rot of soybean. I. Laboratory test for resistance. II. Implications of conidia in epidomology. III. Anti microbial activity of toxins. M. Sc. Thesis, Kansas State University, Manhattan, p 59 Pearson CAS, Leslie JF, Schwenk FW (1986) Variable resistance in Macrophomina phaseolina from corn, soybean and soil. Phytopathology 76:646–649 Pecina-Quintero V, Alvarado MJ, Williams-Alanis H, Almaraz RT, Vandemark GJ (2001) Detection of double stranded RNA in Macrophomina phaseolina. Mycologia 92:900–907 Petrak F (1923) Kykologische Notizen VI. Anals Mycologicity 21:314–315 Pun KB, Sabitha D, Valluvaparidasan V (1998) Studies on seed-borne nature of Macrophomina phaseolina in okra. Plant Dis Res 13:162–164 Punithalingam E (1982) Conidiation and appendage formation in Macrophomina phaseolina (Tassi) Goid. Nova Hedwigia 36:249–290 Purkayastha S, Kaur B, Dilbaghi N, Chaudhary A (2006) Characterization of Macrophomina phaseolina, the charcoal rot pathogen of cluster bean, using conventional techniques and PCRbased molecular markers. Plant Pathol 55:106–116 Reichert I, Hellinger E (1947) On the occurrence, morphology and parasitism of Sclerotium bataticola. Palest J Bot 6:107–147 Reyes-Franco MC, Hernandez-Delgado S, Beas-Fernandez R, Medina-Fernandez M, Simpson J, Mayek-Pérez N (2006) Pathogenic and genetic variability within Macrophomina phaseolina from Mexico and other countries. J Phytopathol 154:447–453 Satto T, Tomioka K, Nakanishi T, Koganezawa H (1999) Charcoal rot of yacon (Smallanthus sonchifolius (Poepp. et Endl.) H.Robinson), Oca (Oxalis tuberosa Molina) and pearl lupin (Tarwi, Lupinus mutabilis Sweet) caused by Macrophomina phaseolina (Tassi) Goid. Bull Shikoku Natl Agri Exp Station 64:1–8 Sharma K, Singh T (2000) Seed and seedling infection of Rhizoctonia bataticola in Vigna radiata. J Mycol Plant Pathol 30:15–18 Siddiqui ZA, Mahmood I (1992) Biological control of root rot disease complex of chickpea caused by Meloidogyne incognita race 3 and Macrophomina phaseolina. Nematologia Mediterranea 20:199–202 Singh D, Nema KG, Singh D (1988) Effect of soil amendment on Rhizoctonia bataticola causing dry root rot of chickpea. International Chickpea Newslett 17:23–25 Singh A, Bhowmik TP, Chaudhary BS (1990) Effect of soil amendment with inorganic and organic sources of nitrogenous manures on the incidence of root rot and seed yield in sesamum. Indian Phytopathol 43:442–443 Singh BP, Saikia R, Yadav M, Singh R, Chauhan VS, Arora DK (2006) Molecular characterization of Fusarium oxysporum f. sp. ciceri causing wilt of chickpea. Afr J Biotechnol 5:497–502 Sobti AK, Bansal RK (1988) Cultural variability among three isolates of R. bataticola from groundnut. Indian Phytopathol 41:149–151 Spencer JFT, Spencer DM (1997) Ecology: where yeasts are. In: Spencer JFT, Spencer DM (eds) Yeasts in natural and artificial habitats. Springer-Verlag, Germany, pp 48–58 Srivastava AK, Arora DK (1997) Evaluation of a polyclonal antibody immunoassay for detection and quantification of Macrophomina phaseolina. Plant Pathol 46:785–794 Srivastava AK, Singh RB (1990) Effect of organic amendment on interaction of Macrophomina phaseolina and Melpidogyne incognita on fresh bean (Phaseolus vulgaris). New Agric 1:99–100 9 Molecular Characterization and Diagnosis of Macrophomina phaseolina 193 Stultz JR, Snoeyenbos-West O, Methe B, Lovley DR, Chandler DP (2001) Application of the 50 fluorogenic exonuclease assay (TaqMan) for quantitative ribosomal DNA and rRNA analysis in sediments. Appl Environ Microbiol 67:2781–2789 Su G, Suh SO, Schneider RW, Russin JS (2001) Host specialization in the charcoal rot fungus, Macrophomina phaseolina. Phytopathology 91:120–126 Suárez ZH, Rosales LC, Rondón A, González MS (1998) Histological alterations on Psidium guajava roots caused by the nematode Meloidogyne incognita race 1 and the fungi Macrophomina phaseolina and Fusarium oxysporum. Fitopatologı́a Venezolana 11:44–47 Sutton BC (1980) The coelomycetes: fungi imperfecti with pycnidia, acervuli and stromata. Commonw Mycol Inst Assoc Appl Biol, Kew, England, p 696 Thakurji (1979) Chemical control of Macrophomina phaseolina root rot of jute. Indian Phytopathol 32:280–281 Upadhayay RK, Arora DK, Dubey OP (2002) Indian Pest Management, vol II. Aditya Books Pvt. Ltd., New Delhi, India, pp 131–147 Vandemark G, Martnez O, Pecina V, Alvardo M de J (2000) Assessment of genetic relationships among isolates of Macrophomina phaseolina using simplified AFLP technique and two different methods of analysis. Mycologia 92:656–664 White TJ, Bruns T, Lee S, Taylor JW (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for Phylogenetics. In: Innis MA, Gelland DH, Sninsky JJ, White TJ (eds) PCR Protocols: a guide to methods and applications. Academic Press, San Diego, pp 315–322 Widjojoatmodjo MN, Borst A, Schukkink RAF, Box ATA, Tacken NMM, Van Gemen B, Verhoef J, Top B, Fluit AC (1999) Nucleic acid sequence-based amplification (NASBA) detection of medically important Candida species. J Microbiol Methods 38:81–90 Chapter 10 Molecular Diagnosis of Ochratoxigenic Fungi Daniele Sartori, Marta Hiromi Taniwaki, Beatriz Iamanaka, and Maria Helena Pelegrinelli Fungaro Abstract Ochratoxin A (OTA) is one of the most abundant food-contaminating mycotoxins. Its presence in several agricultural commodities has been considered a problem worldwide. This toxin is mainly produced by two genera of fungi: Aspergillus and Penicillium. Ochratoxin A has nephrotoxic, immunosuppressive, and carcinogenic effects; consequently, contamination with OTA presents a major risk for human and animal health. Over the last 5 years, several studies have developed PCR-based assays for identifying and quantifying OTA-producing fungi in food samples. The main objective of these assays is to allow the detection of microorganisms capable of producing OTA, prior to ochratoxin production and accumulation. Several of these attempts will be reviewed and discussed in this chapter. 10.1 Introduction Filamentous fungi can produce a vast variety of secondary metabolites and are rich in genes that encode proteins involved in their biosynthesis (Khaldi et al. 2008). In contrast to the primary metabolites that are common and of vital importance for all living organisms, secondary metabolites are not necessary for survival and cellular differentiation, and their synthesis is often limited to a single family, genus, species, or even strain of fungus (Bennett and Ciegler 1983). The secondary metabolites include mycotoxins, which are small organic molecules with diverse chemical structures and biological activities. Mycotoxins are toxic compounds that are occasionally very hazardous to animals and humans. The main source of mycotoxin exposure is from consuming plant foods, which can be contaminated during harvesting, transport, storage or manufacture, or even in the field (Smith and D. Sartori, M.H. Taniwaki, B. Iamanaka, and M.H.P. Fungaro Centro de Ciências Biológicas, Departamento de Biologia Geral, Universidade Estadual de Londrina, Caixa Postal 6001, CEP 86051-970, Londrina-Paraná, Brazil e-mail: fungaro@uel.br Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_10, # Springer-Verlag Berlin Heidelberg 2010 195 196 D. Sartori et al. O C CH2 CH OH O N C OH O O H H CH3 CI Fig. 10.1 The chemical structure of ochratoxin A Henderson 1991). There is a risk to human health not only through the intake of contaminated foods of vegetal source, but also through foods of animal origin due to mycotoxin-contaminated animal feed ingested by animals. Due to the different molecular structures of these mycotoxins, their influences on human and animal health have a variety of effects; they may be neurotoxic, teratogenic, immunosuppressive, nephrotoxic, hepatotoxic, or carcinogenic (Geisen 1998). About 20 different mycotoxins are significant to human health (Geisen 1998; Bennett and Klich 2003), and demand exists for rapid and reliable techniques to detect mycotoxins and mycotoxin producers (Russell and Paterson 2006). Ochratoxins are a class of mycotoxins produced by some fungal species. There are three types of ochratoxin — A, B, and C — but ochratoxin A (7-(L-b-phenylalanyl-carbonyl)-carboxyl-5-chloro-8-hydroxy-3,4-dihydro-3R-methyl isocumarin, or OTA; Fig. 10.1) is the most toxigenic and is one of the most common food mycotoxins. Although this toxin was described many years ago, it has only recently begun to receive considerable attention (van der Merwe et al. 1965). As reviewed by Petzinger and Weidenbach (2002), OTA-contaminated foods are abundant in several countries. About 57% of approximately 6,500 food samples examined in Europe contained amounts of OTA above the detection limit of 0.01 mg kg 1 (Wolff et al. 2000). The most important examples are grains, coffee beans, spices, nuts, grapes, and figs (Bayman et al. 2002; Jorgensen and Jacobsen 2002; Battilani et al. 2003; Taniwaki et al. 2003). OTA is not totally decomposed during most food processing steps such as cooking, washing, and fermenting. Because of this, OTA has also been detected in manufactured food products such as bread, beer, wine, coffee, and chocolate (Jorgensen 1998; Visconti et al. 2001). Kidneys are the main target organs of OTA. Nephrotoxic effects have been demonstrated in all the mammalian species tested so far. For humans, there is abundant circumstantial evidence connecting OTA ingestion with severe nephropathy (Mantle and McHugh 1993; Maaroufi et al. 1995; Wafa et al. 1998). The carcinogenicity of OTA in rodents is also well established, although in humans a correlation between cancer and exposure to OTA has not been proven directly. Based on these facts, the International Agency for Research on Cancer (IARC 1993) classified OTA as a possible human carcinogen (group 2B) in 1993. More 10 Molecular Diagnosis of Ochratoxigenic Fungi 197 recent studies have documented significant effects of OTA on immune response. According to Petzinger and Weidenbach (2002), these additional effects have gained increased attention since it was observed that they may occur even at very low concentrations of OTA. Based on the risk that OTA presents to human health, the European Union has imposed regulations on maximum levels of OTA in cereals and cereal products, dried vine fruit, and all products derived from these items, as well as roasted and soluble coffee. For dried vine fruits (raisins, currants, and sultanas), the maximum tolerable level of OTA is set at 10 mg kg 1; the maximum level permitted in wines and grapes is 2 mg kg 1. For raw cereal grains and all cereal-derived products for direct human consumption, the maximum tolerable levels of OTA are 5 mg kg 1 and 3 mg kg 1, respectively. Finally, the maximum level of OTA in coffee is 5 mg kg 1 for both roasted coffee beans and ground coffee and 10 mg kg 1 for instant coffee (Official Journal of The European Union, 2005). The European Union is at present considering whether a maximum level for OTA should be established in other foods, such as cocoa. A particular species of fungus may produce several mycotoxins, but not all fungi produce mycotoxins. In addition, the type and quantity of a certain mycotoxin are associated with specific strains, while environmental conditions and available nutrients are determinants for the growth of the fungus and its mycotoxin production. The economically most important OTA producers were recently reviewed by Varga and Kozakiewicz (2006): Penicillium verrucosum in cereals (Lund and Frisvad 2003); Aspergillus niger and A. carbonarius in grapes (Battilani and Pietri 2002); Aspergillus ochraceus, A. niger and A. carbonarius in coffee (Bucheli and Taniwaki 2002; Taniwaki et al. 2003), and Aspergillus alliaceus in figs (Bayman et al. 2002). However, recent advances in molecular biology and fungal metabolite analysis resulted in the description of some important new OTA-producing species by European researchers (Frisvad et al. 2004). Aspergillus westerdijkiae, which closely resembles A. ochraceus, is now recognized as the main OTA producer in coffee. The severe consequences of OTA contamination demand efficient and costeffective methodologies for detecting OTA producers in food. In this chapter, we present some of the relevant molecular approaches that have been used to detect and quantify the main OTA producers in food. 10.1.1 Molecular Markers for the Detection of Ochratoxigenic Fungi The traditional schemes for the isolation and identification of ochratoxigenic fungi from food samples are time-consuming and require a high knowledge of fungal taxonomy. Even with taxonomic expertise, identification is commonly difficult in some genera of fungi that contain a large number of closely related species. The application of molecular biology techniques can help to overcome these problems 198 D. Sartori et al. because it can reduce the time for identification from days to hours and also allow precise species identification. Polymerase chain reaction (PCR) is a technique that was developed in 1985 for the in vitro amplification of specific segments of DNA (Saiki et al. 1985; Mullis and Faloona 1987). This technique has allowed the precise identification and fast detection of ochratoxigenic species without the need for isolating pure cultures. The selection of target sequence specific for a given mycotoxin-producing fungus is a key process in the development of a PCR-based diagnostic assay. These target sequences used for designing PCR primers may be didactically divided into two groups: (a) anonymous DNA sequences and (b) functional genes (Carter and Vetrie 2004). Anonymous DNA sequences are obtained from an unbiased sample of genomic DNA and may or may not contain functional genes. Developing markers from anonymous sequences requires comparative analyses between the DNA profiles of related species generated from randomly amplified fragments by random amplified polymorphic DNA (RAPD) or amplified fragment length polymorphism (AFLP) (Williams et al. 1990; Vos et al. 1995). Both methodologies have proven to be powerful taxonomic instruments, especially at low taxonomic positions. The amplification patterns produced by RAPD and AFLP analysis allow discrimination between species and distinct isolates of a single species. Polymorphisms are recognized by the presence or absence of amplified fragments at each RAPD or AFLP locus. DNA bands that are exclusively present in all isolates of a certain toxigenic species may be cloned and sequenced. Once an RAPD or AFLP marker is sequenced, it can be converted into a robust PCR-based marker. Thus, RAPD and AFLP have been applied successfully for revealing specific marker sequences (Schmidt et al. 2003, 2004b; Fungaro et al. 2004a; Sartori et al. 2006). Such sequences have been used to design species-specific primers that allow the identification and detection of some ochratoxigenic species in food samples. As mentioned above, the sequences of functional genes may also be used as targets for PCR primers to detect mycotoxigenic fungi. However, in contrast to other mycotoxins, the OTA biosynthetic pathway has not been well characterized in any of the OTA-producing species; consequently, the genes that encode enzymes involved in the biosynthesis of this secondary metabolite are poorly known. Because of this, the several PCR-based assays developed during the last 5 years have used genes that were not associated to mycotoxin biosynthesis: ribosomal RNA, b-tubulin, and calmodulin genes (Perrone et al. 2004; Patiño et al. 2005; Morello et al. 2007). 10.1.2 PCR-Detection and Quantification of Ochratoxigenic Species with Sequences Not Associated to Mycotoxin Biosynthesis Ochratoxin A was discovered as a secondary metabolite of A. ochraceus strains, which belongs to Aspergillus section Circumdati. Based on a polyphasic taxonomy, 10 Molecular Diagnosis of Ochratoxigenic Fungi 199 which takes into account all accessible phenotypic and genotypic data and integrates them in a consensus classification, 20 species are distinguished into Aspergillus section Circumdati. The taxonomy of this section remains in progress, and Frisvad et al. (2004) recently proposed the division of A. ochraceus into two species, A. ochraceus and A. westerdijkiae. Several species in the section Circumdati are able to produce OTA in culture medium, but the main culprit species for the presence of OTA in food are A. ochraceus and A. westerdijkiae (Frisvad et al. 2004). A. westerdijkiae and A. ochraceus are very similar, and several isolates previously identified as A. ochraceus are now recognized as A. westerdijkiae, including the original OTA-producing strain (NRRL 3174). Amplification and sequencing of the ITS1-5.8S-ITS2 region from several Brazilian strains of both species showed specific nucleotide variations that distinguish A. westerdijkiae and A. ochraceus (Fungaro et al. 2004b). In ITS1, all sequences of A. westerdijkiae differed from the A. ochraceus sequences by possessing a C instead of a T at positions 76 and 80. In addition, A. ochraceus has a deletion of a T at position 89. In ITS2, specific nucleotides at position 494–495 (AT) characterized the strains of A. westerdijkiae, compared to a TC at this position in A. ochraceus. Moreover, a T at position 487 is deleted only in A. ochraceus strains. Similarly, Morello et al. (2007) detected 39 species-specific single nucleotide polymorphisms within the b-tubulin genes from A. westerdijkiae and A. ochraceus, most of them (97.4%) in intronic regions. Eleven nucleotide substitutions and one heptanucleotide insertion/deletion were found in intron 3 (107 bp). Intron 4 (103 bp) was found to have six substitutions, one pentanucleotide insertion/deletion, and one dinucleotide insertion/deletion. Seven substitutions were found in intron 5 (87 bp). The first report of a diagnostic PCR assay for A. ochraceus occurred in 2003 (Schmidt et al. 2003). The authors investigated the genetic relatedness among 70 strains with AFLP markers. A certain number of AFLP bands characteristic for A. ochraceus were detected. Three of these bands were cloned and sequenced, after which the sequences were used to design three primer pairs specific for A. ochraceus. The specificity of the primer pair OCA-V/OCA-R (Table 10.1) was tested with DNA of several different target strains as well as the closely related Aspergillus and Penicillium spp. and DNA isolated from noninfected green coffee. However, this primer pair is able to amplify DNA sequence from both A. ochraceus and A. westerdijkiae because it was developed previous to the division of formal A. ochraceus species into the two species mentioned above. Patiño et al. (2005) developed a specific PCR assay (OCRA1/OCRA2; Table 10.1) for the detection of A. ochraceus on the basis of ITS sequence comparison between several strains of Aspergillus species. The specificity of the primer pair was tested on a number of Aspergillus, Penicillium, Cladosporium, Botrytis, and Alternaria strains commonly associated with grapes, cereals, and coffee. A single fragment of about 400 bp was only amplified from the genomic DNA of A. ochraceus strains. No product was amplified from genomic DNA from Aspergillus isolates other than A. ochraceus or from other genera. According to the authors, the sensitivity of the PCR assay based on ITS sequences was higher (1 and 10 pg of DNA template per reaction) than one based on a single copy gene (0.1 and Reference Schmidt et al. (2003) 200 Patiño et al. (2005) Dao et al. (2005) Morello et al. (2007) Sartori et al. (2006) Patiño et al. (2005) Dobson and O’Callaghan (2004) Mulè et al. (2006) Schmidt et al. (2004a, b) Atoui et al. (2007) Fungaro et al. (2004a) Selma et al. (2008) Perrone et al. (2004) Dobson and O’Callaghan (2004) Bogs et al. (2006) Geisen et al. (2004) D. Sartori et al. Table 10.1 Primer sequences used for the detection of ochratoxin-A producing fungi Species Utility Target region Primer pair A. ochraceus Species-specific detection AFLP marker F50 ATACCACCGGGTCTAATGCA R50 TGCCGACAGACCGAGTGGATT A. ochraceus Species-specific detection rRNA gene F50 CTTCCTTAGGGGTGGCACAGC R50 GTTGCTTTTCAGCGTCGGCC A. ochraceus Species-specific detection pks gene F50 CATCCTGCCGCAACGCTCTATCTTTC R50 CAATCACCCGAGGTCCAAGAGCCTCG A. westerdijkiae Species-specific detection b-tubulin gene F50 TGATACCTTGGCGCTTGTGACG R50 CGGAAGCCTAAAAAATGAAGAG and quantification A. niger Species-specific detection RAPD marker F50 CAGTCGTCCAGTACCCTAAC R50 GAGCGAGGCTGATCTAAGTG A. carbonarius Species-specific detection rRNA gene F50 GCATCTCTGCCCCTCGG R50 GGTTGGAGTTGTCGGCAG A. carbonarius Species-specific detection pks gene F50 TGGGTATGCGCGGGGTGAGGGTAT R50 CCGTAGGCTTCGAAAAACTGACAC A. carbonarius Specie-specific detection cmdA gene F50 CCGATG GAGGTCATGACATGA R50 AATGCGAACCGGATATTAACTTCTG and quantification A. carbonarius Species-specific detection AFLP marker F50 GAATTCACCACACATCATAGC R50 TTA ACTAGGATTTGGCATTGA AC A. carbonarius Specie-specific detection pks gene F50 AATATATCGACTATCTGGACGAGCG and quantification R50 CCCTCTAGCGTCTCCCGAAG A. carbonarius Species-specific detection RAPD marker F50 AGGCTAATGTTGATAACGGATGAT R50 GCTGTCAGTATTGGACCTTAGAG A. carbonarius Species-specific detection pks gene F50 CCCTGATCCTCGTATGATAGCG-30 R50 CCGGCCTTAGATTTCTCTCACC-30 A. carbonarius Species-specific detection cmdA gene F50 AAGCGAATCGATAGTCCACAAGAATAC R50 TCTGGCAGAAGTTAATATCCGGTT P. verrucosum Species-specific detection pks gene F50 TGCACGACCGGGACAACATCA R50 CCGTAGGCCTCCACAAAATCTG P. nordicum Species-specific detection nrps gene F50 AGTCTTCGCTGGGTGCTTCC R50 CAGCACTTTTCCCTCCATCTATCC P. nordicum Species-specific detection pks gene F50 TACGGCCATCTTGAGCAACGGCACTGCC R50 ATGCCTTTCTGGGTCCGATA 10 Molecular Diagnosis of Ochratoxigenic Fungi 201 1 ng of DNA template per reaction). The authors did not mention the new species A. westerdijkiae, and the primer pair presumably does not distinguish between A. westerdijkiae and A. ochraceus. Morello et al. (2007) further exploited the genetic variation found between the b-tubulin gene sequences obtained from A. ochraceus and A. westerdijkiae with the aim of developing primers specific for A. westerdijkiae. The primer pair Bt2Aw-F/ Bt2Aw-R was designed to specifically amplify A. westerdijkiae (Table 10.1). A 347-bp amplicon was visualized in all A. westerdijkiae isolates, but no PCR product was observed in A. ochraceus isolates. The Bt2Aw primers were successfully applied in detecting the 347-bp amplicon when using DNA collected from coffee beans inoculated with A. westerdijkiae. The ochratoxigenic species Aspergillus carbonarius and A. niger belong to section Nigri, which is an important group of species in food mycology. As discussed in Samson et al. (2007), black aspergilli are one of the more complex groups in terms of classification and identification, and numerous taxonomic schemes have been proposed. The differences between some species belonging to section Nigri are very slight and their discrimination requires molecular analysis. A total of 16 species are recognized in Aspergillus section Nigri: A. aculeatus, A. brasiliensis, A. carbonarius, A. costaricaensis, A. ellipticus, A. ellipsoides, A. japonicus, A. foetidus, A. homomorphus, A. heteromorphus, A. lacticoffeatus, A. niger, A. piperis, A. sclerotioniger, A. tubingensis, and A. vadensis, with the latter taxon recently described as a new species (Samson et al. 2004; de Vries et al. 2005). A. niger sensu stricto, A. tubingensis, A. foetidus and A. brasiliensis are morphologically identical and collectively have been called the A. niger aggregate (Parenicová et al. 2001). Although the taxa included in the A. niger aggregate are morphologically indistinguishable, they differ in their ability to produce OTA and other metabolites. The ability of species other than A. niger sensu stricto within A. niger aggregate to produce OTA remain uncertain, probably due to the complexity of species identification. PCR-restriction fragment length polymorphism (RFLP) analysis of the ITS15.8S-ITS2 region allows the four A. niger aggregate taxa to be classified in two patterns (N and T). A. foetidus and A. tubingensis are classified as type T, and A. niger and A. brasiliensis are classified as type N by the Cabañes group (Accensi et al. 1999; Accensi et al. 2001). According to these authors, all OTA-producing strains were classified as pattern N, while none of the pattern T isolates produced OTA. Ueno et al. (1991) described an A. foetidus isolate that is able to produce OTA. However, according to Samson et al. (2004), no strains of A. foetidus sensu stricto produce OTA. Consistent with this analysis, the strain CBS 618.78 of A. foetidus that was described as an OTA producer was later shown to be A. niger and not A. foetidus (Samson et al. 2004). Although it was assumed for several years that A. tubingensis was not able to produce OTA (Samson et al. 2004), two research groups recently found OTAproducing isolates of this species (Medina et al. 2005; Perrone et al. 2006). We have analyzed several isolates within the A. niger aggregate (obtained from Brazilian coffee beans and dried fruit from worldwide origin) in our own laboratory and 202 D. Sartori et al. found that only A. niger sensu stricto was an OTA producer; i.e., none of the A. tubingensis or A. foetidus isolates analyzed was able to produce OTA (unpublished data). This situation demonstrates the importance of the development of specific markers for the identification and detection of a particular ochratoxigenic fungal species. A specific PCR assay for the detection of A. carbonarius was developed by Patiño et al. (2005) based on ITS sequences. The primer pair CAR1/CAR2 generated an amplicon of 420 bp exclusively from A. carbonarius genomic DNA (Table 10.1). Schmidt et al. (2004b) used AFLP to detect specific markers for A. carbonarius. A certain number of amplified fragments were found to be specific to this species. The marker fragments were cloned, sequenced, and used to design a specific primer pair to detect this species. The primer pairs A1B-fw/A1B-rv and C1B-fw/C1B-rv amplify 189 bp and 351 bp fragments, respectively, in all A. carbonarius isolates tested (Table 10.1). Based on an alignment of calmodulin (cmdA) gene sequences, Perrone et al. (2004) identified regions suitable to design specific PCR primers for the detection of A. carbonarius. The primer pair CARBO1/ 2 produced a PCR product of 371 bp with a sensitivity of about 12 pg when using pure total genomic DNA. Although the PCR assay was useful in screening isolates of black aspergilli from grapes, the authors did not use it to detect A. carbonarius strains directly from sample materials. Several strains representing closely related black aspergilli, i.e., A. carbonarius, A. niger, and A. tubingensis, were analyzed by RAPD with the aim of developing species-specific primers for the detection of A. carbonarius in coffee beans (Fungaro et al. 2004a). A typical RAPD pattern is shown in Fig. 10.2. Some Fig. 10.2 Amplification of polymorphic DNA from A. niger (lanes 1–6), A. tubingensis (lanes 7–18), and A. carbonarius (lanes 19–24) strains with the OPX7 random primer. The molecular weight standard (M) is a 1-kb DNA ladder 10 Molecular Diagnosis of Ochratoxigenic Fungi 203 DNA bands were present in all A. carbonarius strains and absent in all strains of A. niger and A. tubingensis. One of these bands was cloned and sequenced, and then used to design a primer pair specific to A. carbonarius (OPX7809-F/OPX7809-R) (Table 10.1). Using this primer-pair, the authors successfully detected an amplicon of 809 bp when DNA from coffee beans infected with A. carbonarius strains was used. No cross-reaction was observed with DNA from coffee beans infected with closely related black aspergilli. Similarly, based on RAPD markers, Sartori et al. (2006) developed specific primers to detect A. niger (Table 10.1). The primer pair denoted OPX7372F/ OPX7372R generated an amplicon of 372 bp in all A. niger stricto sensu isolates, and no amplification product was observed in reactions using DNA from related species. This PCR assay was successfully applied in detecting A. niger in coffee beans. Brazil is the largest coffee bean producer and exporter in the world. Studies concerning fungi with the potential for colonizing Brazilian coffee beans and producing OTA showed that A. ochraceus (now A. westerdijkiae), A. carbonarius, and A. niger are the major species in Brazilian coffee beans. Based on this observation, our group developed a multiplex PCR assay that can detect these three target fungi species directly from coffee bean samples (Sartori et al. 2006). Multiplex PCR (m-PCR) is a procedure that allows the simultaneous amplification of more than one target sequence in a single PCR reaction, decreasing the number of reactions that must be performed to assess the possible presence of different species in a food sample. Sartori et al. (2006) first analyzed the value of the m-PCR assay with DNA obtained from coffee beans inoculated with these three species. Figure 10.3a shows the amplification profiles from simultaneous use of the primer pairs designed for A. westerdijkiae, A. carbonarius, and A. niger. Amplification products of 260, 809, and 372 bp in a single PCR reaction confirmed the presence of Fig. 10.3 (a) Amplification products obtained from DNA isolated from inoculated coffee beans. Lane 1: DNA from coffee beans inoculated with A. ochraceus amplified with OCA V and OCA R primers; lane 2: DNA from coffee beans inoculated with A. niger amplified with OPX7F372 and OPX7R372 primers; lane 3: DNA from coffee beans inoculated with A. carbonarius amplified with OPX7F809 and OPX7R809 primers; lane 4: multiplex PCR using DNA from coffee beans inoculated with A. ochraceus, A. niger, and A. carbonarius amplified with all three sets of primer pairs; lane 5: negative control (DNA from coffee beans without fungal inoculation); lane 6: positive PCR control (A. niger DNA amplified with the primers ITS1 and ITS4). (b) Multiplex PCR obtained from naturally contaminated coffee beans. Lane 1: detection of A. niger; lane 2: no fungi detected; lane 3: detection of A. niger; lane 4: detection of A. niger and A. ochraceus; lane 5: no fungi detected; lane 6: detection of A. niger and A. ochraceus; lane 7: positive control of the multiplex assay (DNA from coffee beans inoculated with A. carbonarius, A. niger and A. ochraceus). Reproduced from Sartori et al. (2006) with permission) 204 D. Sartori et al. A. westerdijkiae, A. carbonarius, and A. niger, respectively. The usefulness of the multiplex PCR assay was also analyzed in coffee bean samples collected on farms. As shown in Fig. 10.3b, this methodology successfully allowed the detection of amplification products from naturally occurring fungi in coffee beans. Penicillium verrucosum has been commonly isolated from cereal crops and is the principal OTA-producing fungus in cool, damp climatic regions (Pitt and Hocking 1997). This species is morphologically very similar to the related species P. nordicum, which is mainly isolated from proteinaceous foods like cheese and fermented meat (Larsen et al. 2001; Castellá et al. 2002). P. nordicum is a high OTA producer in vitro, but until now the ability of this species to produce OTA in its natural environment has not been tested (Bogs et al. 2006). Castellá et al. (2002) used RAPD, AFLP, and ITS sequencing to characterize two groups of Penicillium OTA-producing strains that differed in their ability to produce OTA, with group I containing mainly high-producing strains, and group II containing moderate to nonproducing strains. The strains from group I originate from foods, such as cheese and meat products, while the strains from group II originate from plants. The ribosomal ITS1-5.8S-ITS2 sequences were similar, except for two single nucleotide exchanges in several strains of each group. Group I was recognized as P. nordicum and Group II as P. verrucosum. The authors did not attempt to design species-specific primers to detect either species of Penicillium. Although conventional PCR is a valuable tool for detecting and monitoring mycotoxigenic fungi, it is not appropriate to quantify a given fungus species in a food sample. Small differences in reaction efficiency per cycle can result in a substantial difference in the final product quantity, and so it is very difficult to extrapolate the initial concentration of the template in the sample from the final product (Hill 1996). Fortunately, the introduction of the real-time PCR technology has increased the reliability of PCR results compared to those obtained by conventional methods, thus opening new avenues for quantifying ochratoxigenic fungi in food. Real-time PCR is more sensitive than classical PCR and does not require gel electrophoresis. The analysis can be concluded in less than 5 h. These attributes of real-time PCR significantly reduce time and manual labor, making it appropriate for large-scale analyses. The use of fluorophores is common to most of these methods and is described in detail by Boysen et al. (2000). By using real-time PCR it is possible to detect an increase in fluorescence emission during the reaction which is proportional to the initial copy number of the target sequence. The initial amount of template DNA is inversely proportional to a parameter measured for each reaction, which is denoted as the threshold cycle (Ct). The Ct value is the PCR cycle when the fluorescence signal increases above the background threshold. The application of this method to natural samples can provide an estimate of infection by a given species. Because A. westerdijkiae consistently produces large amounts of OTA, Morello et al. (2007) evaluated the potential of the real-time PCR approach for quantification of this species in coffee beans (Fig. 10.4). A real-time PCR standard curve was obtained with a range of initial amounts of A. westerdijkiae total DNA (20; 10; 5; 1; 0.5 and 0.1 ng per reaction) showing a good correlation (r2 ¼ 0.982). Green 10 Molecular Diagnosis of Ochratoxigenic Fungi 205 1.2 1.1 A westerdijkiae 142 1 A ochraceus 14A 0.9 Blank Fluorescence 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 5 10 15 20 25 Cycle 30 35 40 4 b Fig. 10.4 Discrimination between A. westerdijkiae and A. ochraceus by (a) conventional PCR and (b) real-time PCR. Reproduced from Morello et al. (2007) with permission coffee beans were inoculated with 106 A. westerdijkiae conidia and incubated for 192 h at 28 C. DNA extraction and a colony forming unit (cfu) assay were performed every 48 h. A high correlation was observed between the cfu data and the fungal DNA content in the coffee beans (Fig. 10.5). The authors also assessed the sensitivity of this method in order to detect A. westerdijkiae in coffee beans. Serial dilutions (10 1–10 9) of DNA extracted from infected coffee beans after 48 h of incubation generated a positive signal at up to 10 5 dilution, showing that less than 10 and more than 1 single copy of the A. westerdijkiae haploid genome can be detected by this methodology. This value also indicated that less than 10 haploid genomes could be detected per 0.1 g of coffee beans. Thus, the real-time PCR assay was more than 100 times more sensitive than the cfu technique. A quantitative real-time PCR assay was developed to detect and quantify A. carbonarius in grapes as a possible tool for predicting potential ochratoxigenic risk (Mulè et al. 2006). The species-specific primers and probes used by the authors were derived from conserved regions of the A. carbonarius calmodulin gene. The quantification of fungal genomic DNA in naturally contaminated grapes was performed using the TaqMan signal versus spectrophotometrically measured DNA quantities (log10) calibration curve with a linearity range from 50 to 5  10 4 ng of DNA. A positive correlation (r2 ¼ 0.92) was found between A. carbonarius DNA content and OTA concentration in naturally contaminated grape samples. D. Sartori et al. 9 9 8 8 7 7 6 6 5 5 4 4 3 Log cfu/coffee bean 2 Log copy number/coffee bean 1 r = 0.859293 0 0 48 95 Time (hours) 144 3 2 1 0 192 Log copy number/coffee bean Log cfu coffee bean 206 Fig. 10.5 Comparison of cfu data and the haploid genome copy number of A. westerdijkiae in inoculated coffee beans. Reproduced from Morello et al. (2007) with permission The sensitivity of the PCR method is crucial in the detection of foodborne microorganisms. Unfortunately, there is no standard for reporting sensitivity. Some authors refer to sensitivity as the minimum picograms of DNA that can be detected (Bluhm et al. 2002; Schmidt et al. 2004a; Patiño et al. 2005), others refer to it as the minimum percentage of infected grains in a sample (Schmidt et al. 2004a), and more recently, the lowest detectable number of haploid genomes was also used (Mulè et al. 2006). To eliminate confusion and uncertainties regarding sensitivity, a single method for sensitivity calculation should be adopted. We suggest that the number of haploid genomes per gram of sample is the most convenient metric with which to indicate PCR sensitivity. 10.1.3 PCR-Based Detection and Quantification of Ochratoxigenic Species with Biosynthetic Pathway Genes Various enzymes can be expected to catalyze key reactions in the formation of OTA based on its structure. Some teams of researchers are currently looking for genes related to OTA biosynthesis (Lebrihi et al. 2003; Geisen et al. 2004; Atoui et al. 2006; O’Callaghan et al. 2006; Bogs et al. 2006). A polyketide synthase is predicted to be involved in OTA biosynthesis because the isocoumarin group of OTA is a pentaketide likely to be formed from acetate and malonate via a polyketide synthesis pathway (O’Callaghan et al. 2003). The diversity of polyketide synthase genes has been investigated in A. carbonarius (Atoui et al. 2006). Two nonconserved sequences in the acyltransferase domain of a polyketide synthase gene, denoted Ac12RL3, were used as a target sequence to specifically detect A. carbonarius by PCR. The primer pair Ac12RL_OTAF/ 10 Molecular Diagnosis of Ochratoxigenic Fungi 207 Ac12RL_OTAR (Table 10.1) generated a 141-bp PCR product in all A. carbonarius isolates studied, while no other species gave a positive result with this PCR primer set (Atoui et al. 2007). This specific primer pair was successfully employed to directly quantify A. carbonarius in grape samples. With the same objective, i.e., to quantify A. carbonarius in grape samples, Atoui et al. (2006) used a specific primer pair (Ac12RL_OTAF/Ac12RL_OTAR) (Table 10.1) that was designed from the acyltransferase (AT) domain of the polyketide synthase sequence (Ac12RL3) to amplify a 141-bp PCR product. Using real-time PCR conjugated with SYBR Green I dye, the authors found a positive correlation (r2 ¼ 0.81) between A. carbonarius DNA content and OTA concentration in 72 grape samples. A real-time PCR system based on the otapksPN sequence was used to monitor the growth and OTA production of P. nordicum in wheat (Geisen et al. 2004). A strong correlation between the copy numbers of the otapksPN gene and cfu was observed. Several analytical methods for the detection of OTA are available, and the level of this mycotoxin can readily be measured very accurately in food. However, this kind of analysis only returns a positive result once the toxins have been formed. Similarly, several methods for the detection of ochratoxigenic species have been described, but the presence of an ochratoxigenic fungus in a food sample does not ultimately indicate the production OTA. The formation of OTA depends strongly on environmental conditions such as substrate, water activity, pH, and temperature. Based on these points, the measurement mycotoxin gene expression would allow more meaningful monitoring of OTA in food; these genes are frequently expressed some days prior to the mycotoxin production and thus would allow an early warning (Schmidt-Heydt and Geisen 2007). According to some authors, the expression analysis of key mycotoxin biosynthetic genes might be useful as Hazard Analysis and Critical Control Point (HACCP) for the food industry (Geisen et al. 2004; Niessen 2007). The first relevant report of the cloning and characterization of putative polyketide synthase gene (pks) from OTA-producing Aspergillus was provided by O’Callaghan et al. (2003). These authors used a molecular strategy denoted “Suppression Subtractive Hybridization PCR-Based.” The predicted amino acid sequence of a 1.4-kb clone shared 28–35% identity with acyltransferase regions from fungal polyketide synthases found in the databases. Based on reverse transcription PCR studies, the authors showed that this pks gene is expressed only under OTA-permissive conditions and only during the early stages of mycotoxin synthesis. A mutant in which the pks gene has been interrupted was not able to synthesize OTA. The authors later examined OTA production by A. ochraceus grown under different nutritional and environmental conditions. Quantifications of pks transcript accumulation showed that pks transcription is tightly linked to OTA production (O’Callaghan et al. 2006). As reviewed by Niessen (2007), the University College Cork (Ireland) filed world wide (WO 2004/072224) as well as European (EP 1592705A2) patent applications based on Irish priority application (IE 20030095) based on O’Callaghan’s results. The patent claims cover the use of the sequence for the purpose of detecting OTA producers as well as its use for primer walking. 208 D. Sartori et al. Geisen et al. (2004) used degenerate primers to detect and characterize a portion of a polyketide synthase gene from Penicillium nordicum. All analyzed P. nordicum strains possessed the fragment, whereas the closely related ochratoxigenic P. verrucosum strains did not. An expression analysis of this gene demonstrated that it is highly induced under OTA-producing conditions but only at low levels under nonproducing conditions. In addition, a strong congruence between otapksPN gene expression and OTA production in wheat was observed. Microarray technology is suitable to analyze gene expression on a global level and may be useful for detecting mycotoxigenic fungi before mycotoxins are produced. For this purpose, the mRNA from a given sample is used to generate a labeled sample, termed the “target,” which is hybridized with a large number of DNA sequences that are immobilized on a solid surface in an ordered array. Schmidt-Heydt and Geisen (2007) developed a microarray (DNA chip) that contains oligonucleotides homologous to genes from several fungal species that are responsible for the biosynthesis of mycotoxins. Consequently, this microarray covers most of the known relevant mycotoxin biosynthesis genes. However, it is important to state that although this gene is really more expressed by a positive strain under OTA-permissive conditions, no information is available about the expression of this gene by OTA-nonproducing strains. A preliminary investigation carried out by our group showed that the pks gene, described by O’Callaghan et al. (2003), is in fact significantly more expressed by A. westerdijkiae when grown in OTA-permissive conditions than when grown in OTArestrictive conditions. However, when we cultivated two negative strains in permissive conditions, the pks gene was expressed at levels similar to those of as a positive strain, even though they did not produce OTA (unpublished data). This probably occurs because other secondary metabolites require the function of the pks gene. This observation stresses the importance of identifying genes that are differentially expressed between OTA-producing strains and OTA-nonproducing strains. 10.2 Conclusions Over the last 5 years, several molecular assays for the identification and fast detection of ochratoxigenic species without the need for isolating pure cultures have been published. These assays include conventional PCR, real-time PCR, RT real-time PCR, and microarray technology. Until now, they have been used in research laboratories to detect putative mycotoxin-producing fungi in culture or even in food samples to obtain information on the epidemiology and ecology of ochratoxigenic species or to acquire basic information on gene expression. The use of molecular assays in routine analyses in the food and feed industries remains a challenge. Specificity, sensitivity and simplicity of analysis are all areas that must be improved before these molecular assays become useful for practical applications. Furthermore, OTA biosynthesis is poorly understood relative to the synthesis pathways of other economically important mycotoxins. Better knowledge of the 10 Molecular Diagnosis of Ochratoxigenic Fungi 209 genes involved in OTA biosynthesis is necessary to effectively predict the risk of OTA production. Even so, we are optimistic that molecular technologies will be useful for large-scale analyses in the near future, and will be regularly used as a preventive approach to minimize ochratoxin entry into the food chain. References Accensi F, Caño J, Figueira L, Abarca ML, Cabañes FJ (1999) New PCR method to differentiate species in the Aspergillus niger aggregate. FEMS Microbiol Lett 180:191–196 Accensi F, Abarca ML, Caño J, Figueira L, Cabañes FJ (2001) Distribution of ochratoxin A producing strains in the A. niger aggregate. Antonie Van Leeuwenhoek 79:365–370 Atoui A, Dao P, Mathieu F, Lebrihi A (2006) Amplification and diversity analysis of ketosynthase domains of putative polyketide synthase genes in Aspergillus ochraceus and Aspergillus carbonarius producers of ochratoxin A. Mol Nutr Food Res 50:448–493 Atoui A, Mathieu F, Lebrihi A (2007) Targeting a polyketide synthase gene for Aspergillus carbonarius quantification and ochratoxin A assessment in grapes using real-time PCR. Int J Food Microbiol 115:313–318 Battilani P, Pietri A (2002) Ochratoxin A in grapes and wine. Eur J Plant Pathol 108:639–643 Battilani P, Pietri A, Bertuzzi T, Languisco L, Giorni P, Kozakiewicz Z (2003) Occurrence of ochratoxin A – producing fungi in grapes grown in Italy. J Food Prot 66:633–636 Bayman P, Baker JL, Doster MA, Michailides TJ, Mahoney NE (2002) Ochratoxin production by the species Aspergillus ochraceus group and Aspergillus alliaceus. Appl Environ Microbiol 68:2326–2329 Bennett JW, Ciegler A (1983) Secondary metabolism and differentiation in fungi. Marcel Dekker, New York Bennett JW, Klich M (2003) Mycotoxins. Clin Microbiol Rev 16:497–516 Bluhm BH, Flaherty JE, Cousin MA, Woloshuk CP (2002) Multiplex polymerase chain reaction assay for the differential detection of trichothecene and fumonisin producing species of Fusarium in corn meal. J Food Prot 65:1955–1961 Bogs C, Battilani P, Geisen R (2006) Development of a molecular detection and differentiation system for ochratoxin A producing Penicillium species and its application to analyse the occurrence of Penicillium nordicum in cured meats. Int J Food Microbiol 107:39–47 Boysen ME, Jacobson J, Schnurer J (2000) Molecular identification of species from the Penicillium roqueforti group associated with spoiled. Appl Environ Microbiol 66:1523–1526 Bucheli P, Taniwaki MH (2002) Research on the origin and the impact of post-harvest handling and manufacturing on the presence of ochratoxin A in coffee. Food Addit Contam 19:655–665 Carter PN, Vetrie D (2004) Applications of genomic microarrays to explore human chromosome structure and function. Hum Mol Genet 13:297–302 Castellá G, Larsen TO, Cabañes FJ, Schmidt H, Alboresi A, Niessen L, Färber P, Geisen R (2002) Molecular characterization of ochratoxin A producing strains of the genus Penicillium. Syst Appl Microbiol 25:74–83 Dao HP, Mathieu F, Lebrihi A (2005) Two primer pairs to detect OTA producers by PCR method. Int J Food Microbiol 104:61–67 Dobson A, O’Callaghan J (2004) Detection of ochratoxin A producing fungi. Patent no. WO 2004/ 072224 A2 Frisvad JC, Frank JM, Houbraken J, Kujipers AFA (2004) New ochratoxin A producing species of Aspergillus section Circumdati. Stud Mycol 50:23–43 Fungaro MHP, Vissotto PC, Sartori D, Vilas-Boas LA, Furlaneto MC, Taniwaki MH (2004a) A molecular method for detection of Aspergillus carbonarius in coffee beans. Curr Microbiol 49:123–127 210 D. Sartori et al. Fungaro MHP, Magnani M, Vilas-Boas LA, Vissotto PC, Furlaneto MC, Vieira MLC, Taniwaki MH (2004b) Genetic relationships among Brazilian strains of Aspergillus ochraceus based on RAPD and ITS sequences. Can J Microbiol 50:985–988 Geisen R (1998) PCR methods for the detection of mycotoxin producing fungi. In: Bridge PD, Arora KK, Reddy CA, Elander RP (eds) Applications of PCR in micology. CAB International, Cambridge, UK, pp 243–266 Geisen R, Mayer Z, Karolewiez A, Färber P (2004) Development of a real time PCR system for detection of Penicillium nordicum and for monitoring ochratoxin A production in foods by targeting the ochratoxin polyketide synthase gene. Syst Appl Microbiol 27:501–507 Hill WE (1996) The polymerase chain reaction: applications for the detection of foodborne pathogens. Crit Rev Food Sci Nutr 36:123–173 IARC (International Agency for Research on Cancer) (1993) Ochratoxin A. Monographs on the evaluation of carcinogenic risks to humans: some naturally occurring substances, food items and constituents, heterocyclic aromatic amines and mycotoxines. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans 56:489–521 Jorgensen K (1998) Survey pork, poultry, coffee, beer and pulses for ochratoxin A. Food Addit Contam 15:550–555 Jorgensen K, Jacobsen JS (2002) Occurrence of ochratoxin A in Danish wheat and rye, 1992– 1999. Food Addit Contam 19:1184–1189 Khaldi N, Collemare J, Lebrun M, Wolfe KH (2008) Evidence for horizontal transfer of a secondary metabolite gene cluster between fungi. Gen Biol 9:R18.1–R18.10 Larsen TO, Svendsen A, Smedsgaard J (2001) Biochemical characterization of ochratoxin A-producing strains of the genus Penicillium. Appl Environ Microbiol 67:3630–3635 Lebrihi A, Mathieu F, Borgida LP, Guyonvarch AM (2003) Method for the detection of ochratoxin A or citrinin-producing fungi. European patent No. EP1329521 Lund F, Frisvad JC (2003) Penicillium verrucosum in wheat and barley indicates presence of ochratoxin A. J Appl Microbiol 95:1117–1123 Maaroufi K, Achour A, Betbeder AM, Hammami M, Ellouz F, Creppy EE, Bacha H (1995) Foodstuffs and human blood contamination by the mycotoxin ochratoxin A: correlation with chronic intersticial nephropathy in Tunisia. Archiv Toxicol 69:552–558 Mantle PG, McHugh KM (1993) Nephrotoxic fungi in foods from nephropathy households in Bulgaria. Mycol Res 97:205–212 Medina A, Mateo R, López-Ocaña L, Valle-Algarra FM, Jimenez M (2005) Study of Spanish grape mycobiota and ochratoxin A production by isolates of Aspergillus tubingensis and other members of Aspergillus section Nigri. Appl Environ Microbiol 71(8):4696–4702 Morello LG, Sartori D, Martinez ALO, Vieira MLC, Taniwaki MH, Fungaro MHP (2007) Detection and quantification of Aspergillus westerdijkiae in coffee beans based on selective amplification of b-tubulin gene by using real-time PCR. Int J Food Microbiol 119:270–276 Mulè G, Susca A, Logrieco A, Stea G, Visconti A (2006) Development of a quantitative real-time PCR assay for the detection of Aspergillus carbonarius in grapes. Int J Food Microbiol 111 (Suppl 1):S28–S34 Mullis KB, Faloona FA (1987) Specific synthesis of DNA in vitro via a polymerase-catalyzed chain reaction. Methods Enzymol 155:335–335 Niessen L (2007) PCR-based diagnosis and quantification of mycotoxin producing fungi. Int J Food Microbiol 119:38–46 O’Callaghan J, Caddick MX, Dobson AD (2003) A polyketide synthase gene required for ochratoxin A biosynthesis in Aspergillus ochraceus. Microbiology 149:3485–3491 O’Callaghan J, Stapleton PC, Dobson AD (2006) Ochratoxin A biosynthetic genes in Aspergillus ochraceus are differentially regulated by pH and nutritional stimuli. Fungal Genet Biol 43:213–221 Parenicová L, Skouboe P, Frisvad J, Samson R, Rossen L, Hoor-Suykerbyuk MT, Visser J (2001) Combined molecular and biochemical approach identifies Aspergillus japonicus and Aspergillus aculeatus as two species. Appl Environ Microbiol 67:521–527 10 Molecular Diagnosis of Ochratoxigenic Fungi 211 Patiño B, González-Salgado A, González-Jaén MT, Vázquez C (2005) PCR detection assays for the ochratoxin producing Aspergillus carbonarius and Aspergillus ochraceus species. Int J Food Microbiol 104:207–214 Perrone G, Susca A, Stea G, Mulè G (2004) PCR assay for identification of Aspergillus carbonarius and Aspergillus japonicus. Eur J Plant Pathol 110:641–649 Perrone G, Mulè G, Susca A, Battilani P, Pietri A, Logrieco A (2006) Ochratoxin A production and amplified fragment length polymorphism analysis of Aspergillus carbonarius, Aspergillus tubingensis and Aspergillus niger strains isolated from grapes in Italy. Appl Environ Microbiol 72:680–685 Petzinger E, Weidenbach A (2002) Mycotoxins in the food chain: the role of ochratoxins. Liv Produc Sci 76:245–250 Pitt JL, Hocking AD (1997) Fungi and food spoilage. Blackie Academic and Professional, London Russell R, Paterson M (2006) Identification and quantification of mycotoxigenic fungi by PCR. Proc Biochem 41:1467–1474 Saiki R, Scharf S, Faloona F, Mullis K, Horn GT, Ehrlich HA, Arnheim M (1985) Enzymatic amplification of b-globin genomic sequences and restriction site analysis for diagnosis of sickle cell anemia. Science 239:487–491 Samson RA, Houbraken JAMP, Kujipers AFA, Frank JM, Frisvad JC (2004) New ochratoxin A or sclerotium producing species in Aspergillus section Nigri. Stud Mycol 50:45–61 Samson RA, Noomin P, Meijer M, Houbraken J, Frisvad JC, Varga J (2007) Diagnostic tools to identify black Aspergilli. Stud Mycol 59:129–145 Sartori D, Furlaneto MC, Martins MK, Paula MRF, Pizzirani-Kleiner A, Taniwaki MH, Fungaro MHP (2006) PCR method for the detection of potential ochratoxin-producing Aspergillus species in coffee beans. Res Microbiol 157:350–354 Schmidt H, Ehrmann M, Vogel RF, Taniwaki MH, Niessen L (2003) Molecular Typing of Aspergillus ochraceus and construction of species specific SCAR-primers based on AFLP. Syst Appl Microbiol 26:138–146 Schmidt H, Bannier M, Vogel RF, Niessen L (2004a) Detection and quantification of Aspergillus ochraceus in green coffee by PCR. Lett Appl Microbiol 38:464–469 Schmidt H, Taniwaki MH, Vogel RF, Niessen L (2004b) Utilization of AFLP markers for PCRbased identification of Aspergillus carbonarius and indication of its presence in green coffee samples. J Appl Microbiol 97:899–909 Schmidt-Heydt M, Geisen R (2007) A microarray for monitoring the production of mycotoxins in foods. Int J Food Microbiol 117:131–140 Selma MV, Martı́nez-Culebras PV, Aznar R (2008) Real-time PCR based procedures for detection and quantification of Aspergillus carbonarius in wine grapes. Int J Food Microbiol 122:126–134 Smith JE, Henderson RS (1991) Mycotoxins and animal foods. CRC Press, London, pp 816–841 Taniwaki MH, Pitt JI, Teixeira AA, Iamanaka BT (2003) The source of ochratoxin A in Brazilian coffee and its formation in relation to processing methods. Int J Food Microbiol 82:173–179 Ueno Y, Kawakura O, Sugiura Y, Horiguchi K, Nakajima M, Yamamoto K, Sato S (1991) Use of monoclonal antibodies enzyme-linked immunosorbent assay and immunoaffinity column chromatography to determine ochratoxin A in porcine sera coffee products and toxin-producing fungi. In: Castegnero M, Plestina R, Dirheimer G, Chernozemsky IN, Bartsch H (eds) Mycotoxins, Endemic Nephropathy and Urinary Tract Tumours, IARC – Sci – Publ. Lyon, France, pp 71–75 Van der Merwe KJ, Steyn PS, Fourie L, Scott DB, Theron JJ (1965) Ochratoxin A, a toxic metabolite produced by Aspergillus ochraceus Wilh. Nature 205:1112–1113 Varga J, Kozakiewicz Z (2006) Ochratoxin A in grapes and grape-derived products. Trends Food Sci Technol 17:72–81 Visconti A, Pascale M, Cewntonze G (2001) Determination of ochratoxin A in wine and beer by immunoaffinity columm cleanup and liquid chromatographic analysis with fluorometric detection: collaborative study. J AOAC Intern 8:1818–1827 212 D. Sartori et al. Vos P, Hogers R, Bleeker M, Reijans M, Van de Lee T, Hornes M, Frijters A, Pot J, Peleman J, Kuiper M, Zabeau M (1995) AFLP: a new technique for DNA fingerprinting. Nucl Acids Res 23:4407–4414 de Vries RP, Frisvad JC, van de Vondervoort PJI, Burgers K, Kujipers AFA, Samson RA, Visser J (2005) Aspergillus vadensis, a new species of the group of black aspergilli. Antonie Van Leeuwenhoek 87:195–203 Wafa EW, Yahya RS, Sobh MA, Eraky I, el-Baz M, el-Gayar HA, Betbeder AM, Creppy EE (1998) Human ochratoxicosis and nephropathy in Egypt: a preliminary study. Hum Experim Toxicol 17:124–129 Williams JGK, Kubelik AR, Livak KJ, Rafalski JA, Tingey SV (1990) DNA polymorphisms amplified by arbitrary primers are useful as genetic markers. Nucleic Acids Res 18:6531–6536 Wolff J, Bresch H, Cholmakov-Bodechtel C, Engel G, Garais M, Majerus P, Rosner H, Scheur R (2000) Ochratoxin A: contamination of foods and consumer exposure. Archiv Lebensmitt 51:81–128 Chapter 11 Molecular Barcoding of Microscopic Fungi with Emphasis on the Mucoralean Genera Mucor and Rhizopus Youssuf Gherbawy, Claudia Kesselboth, Hesham Elhariry, and Kerstin Hoffmann Abstract A broad range of fungi were isolated from different geographic regions and substrates and identified according to traditional and modern methods. A total of 120 different isolates were assigned to the phyla, Basidiomycota with 8 isolates, Ascomycota with 75 isolates, and “Zygomycota” with 37 isolates. Although morphological characters were able to differentiate the isolates to their phyla and in most cases to the correct genera, a combination of several methods is always recommended because characterization and identification of unknown fungal isolates is highly error-prone if relying on single methods. Sequence-based identification turned out to be reliable for most Ascomycetes and Zygomycetes. But with the ongoing questionable trend to rely on sequences as first source information for species separation, the most serious problems are the annotation problems in public reference databases, the inconsistency of described taxa, and the available reference data. 11.1 Introduction and Background The characterization and identification of organisms is fundamental in biological life sciences. Each individual in general is regarded to be composed of numerous, if not countless, characters. Every definable character could be used for descriptive and comparative studies concerning all applied aspects of life. Such characters could be features of morphology, biochemical composition, physiological characters, Y. Gherbawy and C. Kesselboth Botany Department, Faculty of Science, South Valley University, 83523 Qena, Egypt K. Hoffmann Institute of Microbiology, School of Biology and Pharmacy, University of Jena, Neugasse 25, 07743 Jena, Germany e-mail: Hoffmann.Kerstin@uni-jena.de H. Elhariry Biological Sciences Department, Faculty of Science, Taif University, P.O. Box 888 Taif, Kingdom of Saudi Arabia Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_11, # Springer-Verlag Berlin Heidelberg 2010 213 214 Y. Gherbawy et al. ecological properties, metabolic characteristics, or molecular features ranging from mere nucleotide and amino acid sequences to structures, functions, and regulation. Molecular data have several advantages over other characters because they are not subjected to the highly subjective eye of an investigator if morphological criteria are investigated. Also, they are independent from environmental or nutritional conditions influencing metabolism, cell composition, or cellular appearance. Moreover, molecular data by means of nucleic acid sequences are quite easy to access compared to chemotaxonomical criteria, which require cost-intensive laboratory equipment, as for performance of gas chromatography–mass spectrometry. A major drawback is that chemotaxonomical markers are prone to external influences and physiological variability and therefore hard to reproduce. But a solely single nucleotide sequence of a certain gene marker is not capable to designate an organism to a species or reveal the actual features displayed by that organism. In the end, a pure sequence is useless information without further data on the organism of its origin, the locality of its isolation, and its preferred substrate. In this respect, the identification of an organism must not depend on single methods; it should rather be supported by a bunch of different criteria. Comparing and analysing all available data gives insight into the biodiversity of the Earth and allows the reconstruction of the evolutionary history of the organism. Well-supported taxonomic relationships are necessary for the precise and reliable classification of new and unknown specimens. Consequently, the reconstruction of phylogenetic relationships works for well-studied species only, if based solely on molecular information without knowledge of morphological, physiological, or ecological features. The exploration of new specimens clearly depends on more than one criterion. Taxonomic systems based on reconstructed phylogenetic relationships are essential for a sustainable organisation of biological information and a deeper understanding of the evolution and species diversification (Fenchel and Finlay 2006; Wheeler 2004). Identification and classification of unknown specimens requires a broad sample of well-defined and described reference specimens for comparison (Meyer and Paulay 2005; Hoffmann et al. 2009b). Storage of available information in public databases, and type strains maintained and accessible from culture collections, as well as a consensus in taxonomical classification through the scientific community, are indispensable for precise assignments of species. At the present time, the trend tends to rely mainly if not exclusively on molecular data to identify biological specimens with the focus on so-called “DNA barcodes”, a term created in 1993 (Arnot et al. 1993). The now well-established Consortium for the Barcode of Life (CBOL, http://barcoding.si.edu) is aimed to coordinate the research on DNA barcodes and to establish global standards with an open-access database about species diversity. The search for the universal barcode marker, which distinguishes all living beings or at least large organismic groups is presently a subject of heavy scientific debate. So far, the most promising barcodes for the identification of animals are sequences of the mitochondrial cytochrome oxidase subunit I (CO1 or cox1) (Hebert et al. 2003). cox1 has been successfully applied in various studies identifying a broad range of taxa (Hebert et al. 2004a, b; Johnson and Cicero 2004; Tavares and Baker 2008). But cox1 shows also some applicational difficulties, for instance 11 Molecular Barcoding of Microscopic Fungi with Emphasis 215 in the differentiation of parapatric species which do not share a common ecological habitat (Moritz and Cicero 2004; Aliabadian et al. 2009). Additional promising sequences for metazoan phylogenetics target the small ribosomal subunit (16S rDNA), the internal transcribed spacer (ITS), and the cytochrome b (cob) (Bradley and Baker 2001; Helbig and Seibold 1999; Lemer et al. 2007; Park et al. 2007; Vences et al. 2005). Although cox1 was also proven to be suitable for the identification of algae (Saunders 2005), it is not useful for land plants because of the high intra– interspecific variability in the evolutionary rates of the mitochondrial DNA. An applicable DNA barcode marker for land plants is still discussed. The nuclear ITS region, the plastid trnHspbA intergenic spacer, and the rbcL gene were suggested by several authors (Chase et al. 2005; Kress et al. 2005; Kress and Erickson 2007; Newmaster et al. 2006). For fungi, cox1 is also proposed and was successfully tested for Penicillium (Seifert et al. 2007). But, introns, which were reported occasionally (Woo et al. 2003), or the lack of mitochondrial genomes arising from missing mitochondria in anaerobic gut fungi (Yarlett et al. 1986), as well as non-vertical inheritance of mitochondrial genes caused by parasexuality-driven hyphal anastomoses over the species barrier, disqualifies cox1 as the universal marker for fungi. However, alternative barcode markers like the widely used internal transcribed spacer (ITS) region of the nuclear ribosomal DNA cluster are in some cases insufficiently variable to reliably separate the species apart (Skouboe et al. 1999). Because of their frequent lack of distinguishable morphological characters, especially fungi necessitate a robust DNA-based identification system. Although molecular identification is well-established for distinctive fungal groups, a standardised protocol which is over-all applicable is still missing. The aim of the present study is to elucidate common problems in barcoding concerning the eukaryotic kingdom fungi with emphasis on Zygomycetes exemplified with the prominent genera Mucor and Rhizopus. Traditionally, four major phyla are distinguished within the kingdom Fungi: Chytridiomycota, Zygomycota, Ascomycota and Basidiomycota. But in recent years, it has become evident that this traditional scheme does not reflect the phylogenetic relationships among fungi, especially within the basal fungal lineages. Although the monophyletic phyla Ascoand Basidiomycota are well-characterised sister groups today and combined to the subkingdom of the Dikarya, the classification of the basal fungal lineages is still in flux (Sugiyama 1998; van de Peer et al. 2000; Berbee and Taylor 2001; James et al. 2006; Hibbett et al. 2007). Major chances in recent years were the establishment of the phyla Glomeromycota (Schüßler et al. 2001), Blastocladiomycota (James et al. 2007), and Neocallimastigomycota (Hibbett et al. 2007) formerly embedded in Zygomycota and Chytridiomycota. Both these phyla harbour several important pathogens of plants, fungi, animals, and man, causing chytridiomycoses or zygomycoses. With a presumed 70% increase of zygomycotic infection diseases between the years 1940 and 2000 (Roden et al. 2005), Zygomycetes are of growing medical importance, especially for patients with immunocompromised systems and diabetes mellitus, or for intravenous drug users (e.g. Greenberg et al. 2004; Metellus et al. 2008; Nucci and Marr 2005; Ribes et al. 2000; Walsh et al. 2004). This increase was notably significant even before the beginning of voriconazole prophylaxis and the treatment of aspergillosis infections in immunocompromised patients 216 Y. Gherbawy et al. (Roden et al. 2005; Rogers 2008; Trifilio et al. 2007). Common zygomycotic infections affect the rhino-orbito-cerebral tract, the respiratory tract, gastrointestinal tract, or skin (Iwen et al. 2007; Roden et al. 2005). Species involved in mycoses belong mainly to the order Mucorales and can be classified to the epidemiologically and clinically important genera Rhizopus, Mucor, Lichtheimia (formerly Absidia), Cunninghamella, Rhizomucor, and Apophysomyces (Diwakar et al. 2007; Iwen et al. 2007; de Hoog et al. 2000; Ribes et al. 2000). Because of the growing significance of infections and their different sensibility to antifungal drugs, precise identification down to species level is indispensable (Bal 2006; Cuenca-Estrella et al. 2006; Singh et al. 2005). Comprehensive studies concerning clinically important fungi revealed Rhizopus arrhizus (formerly R. oryzae), R. microsporus, and Mucor circinelloides as the most frequent agents of mucormycoses (Alastruey-Izquierdo et al. 2009; Alvarez et al. 2009). Each of these species has a different susceptibility to antifungals (Alastruey-Izquierdo et al. 2009; Almyroudis et al. 2007; Dannaoui et al. 2003). The possibility to identify clinically important Zygomycetes based on DNA markers was successfully demonstrated in recent years (Iwen et al. 2005, 2007; O’Donnell et al. 2001; Schwarz et al. 2006; Voigt et al. 1999; Voigt and Wöstemeyer 2001; White et al. 2006). We want to show the synergistic supplementation of different easy-to-access data aiming at the identification of fungi with emphasis on the mucoralean genera Mucor and Rhizopus. Although DNA markers usually allow an easier (in theory) distinction of clinically important taxonomic groups, morphology is still essentially required for supporting the species designations and the description of newly identified species (Alastruey-Izquierdo et al. 2010; Hoffmann et al. 2007, 2009b). 11.2 Methodical Section 11.2.1 Isolation, Cultivation and Maintenance of Strains A considerable number of soil and air-borne fungi were isolated from Saudi Arabian fruits and soil, Germany, and Austria (Tables 11.1–11.4). Fungal isolates were cultivated for maintenance and isolation of genomic DNA was done on MEX solid media (30 gL 1 malt extract supplemented with 5 gL 1 yeast extract and 20 gL 1 agar). All strains with a FSU number are deposited at the Fungal Reference Centre, University of Jena, and available upon request (www.prz.uni-jena.de). 11.2.2 Morphological Identification Morphological identification was performed in following the guidelines of description keys commonly used for species identification and the therein recommended 11 Molecular Barcoding of Microscopic Fungi with Emphasis 217 Table 11.1 Basidiomycetes, isolated from Saudi Arabian fruits and soil. Because of lacking morphological traits, a secure identification was not possible. Sequence BLAST results of ITS sequences are given, but were not useful for correct species assignments. Taxonomical affiliations for the best BLAST hits are indicated according to the taxonomy at NCBI. All BLAST results belong to: Basidiomycota, Agaricomycotina, Agaricomycetes FSU-no. GenBank ITS sequence BLAST Taxonomical affiliations Acc.no. BLAST results identity (%) 6258/ 6280 GQ221186 Bjerkandera adusta 98–100 Polyporales, Coriolaceae Thanatephorus 98–100 Cantharellales, GQ221187 cucumeris Ceratobasidiaceae 6263/ GQ221188 Phlebia radiata 98–100 Corticiales, Corticiaceae 6282 GQ221189 Sistotrema 99 Corticiales, Corticiaceae 6301/ 6404 GQ221190 brinkmannii GQ221191 Lactarius 99 Russulales, Russulaceae chrysorrheus Coprinopsis 99 Agaricales, cothurnata Psathyrellaceae Merulius 97–99 Corticiales, Corticiaceae 6291 GQ221192 tremellosus Trametes versicolor 98–99 Polyporales, Coriolaceae 6418 GQ221193 T. ochracea 99 Polyporales, Coriolaceae Tricholoma 99 Agaricales, robustum Tricholomataceae Phellinus igniarius 98 Hymenochaetales, Hymenochaetaceae media and methods. For the identification of Zygomycetes the following keys were used: Benjamin (1979), Ellis and Hesseltine (1965, 1966), Hesseltine and Ellis (1961, 1964, 1966), Zycha et al. (1969), Schipper (1973, 1975, 1976, 1984, 1990), Schipper and Stalpers (1984), and Alastruey-Izquierdo et al. (2010). Species of the phyla Asco- and Basidiomycota were identified using Samson and Frisvad (2004), Raper and Fennell (1965), Raper and Thom (1949), Leslie and Summerell (2006), and Wollenweber and Reinking (1935). Some prominent morphological features for differentiation are summarised in Figs. 11.1–11.3. 11.2.3 Extraction of Genomic DNA and PCR Amplification Extraction of genomic DNA and PCR amplification of marker genes were done as described elsewhere (Einax and Voigt 2003). The primers for amplification were ITS1/4 for the internal transcribed spacer regions 1 and 2 (White et al. 1990), cmd5/6 for calmodulin fragments (Hong et al. 2005), and bt2a/b for beta-tubulin fragments (Glass and Donaldson 1995). The amplicons were purified using the adsorption to glass particles described by Vogelstein and Gillespie (1979). Purified fragments were subjected to direct sequencing using the PCR primers as sequencing primers. (6276, 6279)h 6408h GQ221109, GQ221163, GQ221088 9292b 9320c GQ221110 GQ221111 Aspergillus niger var.niger Aspergillus sclerotiorum Aspergillus ustus Aureobasidium sp. Bipolaris sp. Botrytis sp. Btub: A. niger var. niger [99%] Cmd: A. niger var. niger, A. awamori [98%] ITS: A. sclerotiorum, A. persii, A. bridgeri [99%] Btub: A. sclerotiorum, A. persii [97–98%] ITS: A. insuetus, A. ustus [99%] Btub: A. insuetus, A. ustus [96–97%] Cmd: A. insuetus, A. ustus [94–96%] ITS: A. pullulans [100%] ITS: B. heveae [99%] A. niger var. niger A. sclerotiorum A. ustus A. pullulans B. heveae B. cinerea Y. Gherbawy et al. 9297b GQ221164–GQ221165, GQ221089– GQ221090 GQ221108, GQ221162 218 Table 11.2 Ascomycetes, isolated in Saudi Arabia. Accepted identities based on considerations combined from morphological data and sequence BLAST results of ITS, beta-tubulin or calmodulin sequences Final FSU-no. GenBank Acc.no. Identification based Identification verified by sequence BLAST, identification [BLAST identity] on morphology Acrostalagmus sp. ITS: A. luteoalbus [99%] A. luteoalbus CK1 GQ221095 GQ221096 –GQ221104 Alternaria sp. ITS (9285=9286=9289= 9306=9308= Alternaria sp. (9285, 9286, 9313=9314=9317=9641=9653 / 9315): 9306, 9308, Alternaria sp. [99%] 9641, 9653)a, 9289b, (9313, 9314, 9315, 9317)c 9295d Aspergillus n.d. A. candidus candidus 9293d GQ221105, GQ221159 Aspergillus sp. ITS: Eurotium chevalieri, A. cristatus, A. chevalieri E. amstelodami, A. ruber [98%] Btub: Eurotium chevalieri [100%] GQ221106, GQ221160 Aspergillus flavus ITS: A. flavus var. oryzae [99%] A. flavus var. 9429f var. oryzae Btub: A. flavus var. oryzae [99%] oryzae GQ221107, GQ221161, GQ221087 Aspergillus ITS (6405): A. fumigatus [100%] A. fumigatus (6264, 6405)x Btub (6264): A. fumigatus [100%] fumigatus Cmd (6264): A. fumigatus [99%] GQ221128 GQ221129 GQ221130–GQ221138,GQ221140, GQ221141,GQ221175–GQ221176 Engyodontium sp. Exserohilum sp. Fusarium sp. GQ221139, GQ221174 Fusarium oxysporum GQ221142 GQ221177 Geotrichum candidum var. citri-aurantii n.d. Paecilomyces sp. Chaetomium sp. GQ221124 GQ221125, GQ221172 GQ221126 GQ221127, GQ221173, GQ221092 Cochliobolus sp. Cochliobolus sp. Chaetomiaceae Emericella quadrilineata 8673g, 9639f 9310a 9299d ITS: F. solani, F. oxysporum [99–100%] Btub: F. solani, F. oxysporum [100%] n.d. ITS: Microsphaeropsis arundinis [100%] Btub: P. variotii [97%] C. globosum Cochliobolus sp. C. spicifer C. kuwaitiensis E.quadrilineata E. sp. E. rostratum Fusarium sp. F. oxysporum G. candidum var. citri-aurantii M. arundinis P. variotii (continued) 219 9303a 9305a (8671, 9302, 9427, 9428, 9643, 9298)a, (9294, 9300)b, 6261e, 9296f, 9642g, (9301, 9304)j 9311d GQ221120–GQ221123, GQ221091 ITS(6277=6296=6297=6300; 6289=6298= 6409=9640=9647): B. cinerea, B. fabae, Sclerotinia sclerotiorum [99%] Btub (6297=9640=9647; 6289=6298= 6409): B. cinerea [99%, 95%]1 ITS (6295=6270=6290=6292): C. globosum [100%] Cmd (6270): C. globosum [99%] ITS: C. kusanoi, Drechslera portulacae [94%] ITS: C. spicifer [100%] Btub: C. sp. [90%] ITS: Corynascus kuwaitiensis [99%] ITS: E. quadrilineata, E. nidulans, Emericella miyajii [100%] Btub: E. miyajii, A. parvathecius, E. quadrilineata [99–100%] Cmd: E. quadrilineata [100%] ITS: E. album [96%] ITS: Exserohilum rostratum [99%] ITS: diverse Fusarium sp. [98–100%] Btub (9298, 9304): diverse Fusarium sp. [94–99%]2 Molecular Barcoding of Microscopic Fungi with Emphasis GQ221112 –GQ221119, GQ221166– GQ221171 11 (6277, 6296, 6297, 6298, 6300, 6409)h, (9640, 9647)f, 6289x (6270, 6290, 6292, 6295)h 9312c 9290e 9674a 9309a Final Identification verified by sequence BLAST, identification [BLAST identity] GQ221143–GQ221146,GQ221178– ITS: P. chrysogenum [100%] P. chrysogenum (6268, 6267, GQ221181 Btub: P. chrysogenum 6406)h, [98–100%] 6265x GQ221147 Penicillium sp. ITS: P. expansum [99%] P. expansum 6269h 6293h, 6294x GQ221148 Penicillium molle ITS: diverse Penicillium sp. [100%]3 P. molle (6278, 6281)h GQ221149–GQ221150, GQ221182– Penicillium sp. ITS: P. commune, P. griseoroseum , P. solitum P. solitum var. var. crustosum, P. italicum [99%] crustosum GQ221183, GQ221093–GQ221094 Btub: P. solitum var. crustosum [98–100%] Cmd: P. solitum var. crustosum, P. hirsutum var. allii [99%] GQ221151 Pestalotiopsis sp. ITS: P. clavispora, P. photiniae [100%] Pestalotiopsis sp. 8674a GQ221152 Cochliobolus sp. ITS: Pseudocochliobolus verruculosus [100%] P. verruculosus 9307a 9646b GQ221153 Cladosporium sp. ITS: Retroconis fusiformis [100%] R. fusiformis (9425, 9426, GQ221154–GQ221156, GQ221184– Trichoderma sp. ITS: diverse Trichoderma sp. [98–100%] Trichoderma sp. 9431)a GQ221185 Btub (9425, 9431): Trichoderma viride [93%] GQ221157–GQ221158 Ulocladium sp. ITS: diverse Ulocladium sp. [99%] Ulocladium sp. 9287b, 9319c 1 No adequate reference sequence in database for Botrytis fabae, but morphological clearly B. cinerea 2All these Fusarium species belong to different species with 1–20% sequence differences based on an ITS1-5.8S rDNA-ITS2 alignment (data not shown) 3 No ITS sequence for Penicillium molle available Substrates of isolation: (a) soil, (b) air, (c) wheat, (d) floor, (e) date palm, (f) guava, (g) apricot, (h) Calotropis procera, (j) banana, (x) unknown Identification based on morphology Penicillium chrysogenum 220 Table 11.2 (continued) FSU-no. GenBank Acc.no. Y. Gherbawy et al. 11 Molecular Barcoding of Microscopic Fungi with Emphasis 221 Table 11.3 Zygomycetes, isolated in Saudi Arabia. Accepted identities based on considerations combined from morphological data and sequence BLAST results of ITS sequences Final Identification verified FSU-no. GenBank Acc. Identification identification by sequence BLAST, based on no. [BLAST identidy] morphology (6254, 9673)a, GQ221194– Actinomucor A. elegans [99%] A. elegans 6256b GQ221196 elegans Mucor (6257, 6259)a, GQ221197– M. circinelloides M. circinelloides 9637c GQ221199 circinelloides [99–100%] f. griseof. griseocyanus cyanus GQ221200 Mucor hiemalis Rhizomucor variabilis, 9654c M. hiemalis M. circinelloides, M. hiemalis, M. racemosus, [98–99%] 9635a GQ221201 Mucor varians Rhizomucor variabilis, M. varians M. circinelloides, M. hiemalis, M. racemosus, [98–99%] GQ221202– Rhizopus R. arrhizus var. arrhizus R. arrhizus var. (6255, 6262, [100%] arrhizus GQ221207 arrhizus var. 9651, arrhizus 9655)a, 9648d (6253, 6266)x Substrates of isolation: (a) Soil, (b) Date palm, (c) Poultry farm soil, (d) apricot, (x) Unknown 11.2.4 The Fungal Subphylum Mucoromycotina Out of ten families of the order Mucorales (Mucoromycotina, “Zygomycota”) species representatives for 26 genera were analysed on the basis of sequences of the 18S rDNA, 28S rDNA, actin (act), and translation elongation factor 1alpha (tef). The sequences were retrieved from Genbank (Table 11.5) and subjected to maximum parsimony, maximum likelihood, Bayesian inference, and distance based phylogenetic reconstructions (Fig. 11.4). 11.2.5 Reconstruction of Multigene Phylogenetic Trees Single alignments were carried out using ClustalX version 1.83 (Higgins and Sharp 1988, 1989; Thompson et al. 1997). The alignment consists of 41 taxa and 3,503 characters (1,215 characters for 18S rDNA, 389 characters for 28S rDNA, 807 characters for act, and 1,092 characters for tef). Bayesian inference with MrBayes v3.0b4 (Huelsenbeck and Ronquist 2001; Ronquist and Huelsenbeck 2003) was initiated from a random starting tree. Two runs with each four chains were 222 Y. Gherbawy et al. Table 11.4 Zygomycetes, isolated from various substrates in Europe. Accepted identities based on considerations combined from morphological data and sequence BLAST results of ITS sequences Identification verified by Identification based FSU-no. GenBank Acc. Origin of isolation Final identification sequence BLAST, [BLAST on morphology (isolated by) no. identity] Cunninghamella FSU6521, 6526 diverse C. echinulata 6510-6516, GQ221208– Waddenmeer, Vlieland, The Cunninghamella sp. echinulata 6521, 6526 GQ221209 Netherlands [89–92%] 6250, 6520, 6523 GQ221210– Lichtheimia L. corymbifera [98–100%] L. corymbifera 6250: human skin, Germany; GQ221211, corymbifera 6520: cow; 6523: dung of GQ221217 pigeon, Innsbruck, Austria 6524 GQ221212 soil, Geisenheim, Germany Mortierella alpina M. alpina [100%] M. alpina M. circinelloides [99%] 6251, 6252, 6518 GQ221218– Mucor circinelloides f. M. circinelloides 6251, 6252: human ear, nail, circinelloides f. circinelloides GQ221219, Germany; 6518: Austria GQ221213 6517 GQ221214 Snail, Austria Mucor fragilis M. fragilis [99%] M. fragilis 6519, 6530 GQ221215, 6519: soil, Geisenheim, Mucor hiemalis M. hiemalis [99–100%] M. hiemalis GQ221220 Germany; 6530: human sole of foot, Germany 6274 GQ221221 Human nail, Germany Mucor plumbeus M. plumbeus [98%] M. plumbeus 6527 Soil, Innsbruck, Austria Umbelopsis isabellina n.d. U. isabellina 6522, 6529 GQ221216, 6522: soil, Martell, Italy; Zygorhynchus moelleri Z. moelleri [99%] Z. moelleri GQ221222 6529: human nail, Germany 11 Molecular Barcoding of Microscopic Fungi with Emphasis 223 Fig. 11.1 Light microscopic images of several easy-for-differentiation characters typically for the fungal phyla. (a) regularly septated mycelium typical for Asco- and Basidiomycetes; (b) clamp of basidiomycetes; (c) Perithecium, the sexual reproductive structure of some Ascomycetes (Chaetomium sp.); (d) Zygospore, the sexual reproductive structure of Zygomycetes. Scale bar: (a–b, d) 20 mm; (c) 100 mm conducted for 5,000,000 generations with samples from every 5,000 generation. After discarding the first 25% of the generated trees (burn-in) the consensus tree was calculated using the halfcompat option. Posterior probabilities (in percent) at the nodes represent node confidence values. The Bayesian inferred tree is presented in Fig. 11.4. Distance analysis with distance measure Jukes-Cantor assuming minimum evolution was done with PAUP* v4.0b10 (Swofford 1998); negative branch lengths were prohibited. Bootstrap supports (BS) (Felsenstein 1985; 50% majority rule) were obtained by 1,000 replicates and Jukes-Cantor distances. In Maximum Parsimony, the starting tree was obtained by stepwise addition of the sequences. The sequences were added on a simple basis and one tree was held at each step. Tree-bisection-reconnection (TBR) was the branch-swapping algorithm. Steepest descent was not in effect. “MulTrees” option was in effect. Two trees were retained. The bootstrap support (BS) was calculated with fast-heuristic search and 1,000 replicates. Maximum Likelihood was also carried out using a heuristic search. The number of substitution types was 2 (HKY85 variant) as was the transition/transversion ratio. Assumed nucleotide frequencies were empirical. A molecular clock was not enforced. The starting branch lengths were obtained by the Rogers–Swofford approximation method (Rogers and Swofford 1998). Using stepwise addition of the sequences and choosing as-is for the addition, the starting tree was obtained. TBR was the branch-swapping algorithm. Steepest descent was not in 224 Y. Gherbawy et al. a c1 b1 b3 b2 c2 Fig. 11.2 Light microscopic images of asexual reproductive structures of zygomycetes. (a) sporangium with endogenous mitospores (Mucor circinelloides); (b1) apophysate sporangium of Rhizopus arrhizus; (b2) remaining columella after spore release of R. arrhizus; (b3) Sporangium of R. arrhizus arising opposite rhizoids; (c) Sporangium and columella of Umbelopsis isabellina, a fungus of the phylogenetic basal family Umbelopsidaceae (Fig. 11.4). Scale bar: (a–b) 20 mm; (c) 10 mm effect. “MulTrees” option was in effect. The bootstrap support (BS) was calculated with fast-heuristic search and 100 replicates. Trees and Bootstrap supports of parsimony, likelihood, and distance analyses supported the topology and branch support from the Bayesian inference analyses, and are therefore not shown. BS values equal or greater than 75% in all analyses are indicated as bold branches in Fig. 11.4. 11.2.6 Analysis of the Internal Transcribed Spacer Regions 1 and 2 Including 5.8S rDNA For several isolates of Rhizopus stolonifer, R. arrhizus, and Mucor circinelloides ITS1-5.8S rDNA-ITS2 sequences were generated within this study and deposited in GenBank as accession numbers AM933543-55, AM937531-2, GQ221197-99, GQ221202-07, GQ221218-19, GQ221213. One sequence for Lichtheimia corymbifera was generated as outgroup taxon (AM937530). The following sequences 11 Molecular Barcoding of Microscopic Fungi with Emphasis c a2 a1 225 e1 b1 b2 e2 d d1 Fig. 11.3 Light microscopic images of diverse asexual reproductive structures and conidia observed in Ascomycetes. (a) phialides arising directly from the conidiogenous cell (a1 Aspergillus fumigatus, a2 Stachybotrys sp.); (b) or phialides arising from metulae (b1 Acrostalagmus luteoalbus, b2 Penicillium sp.; c) polyblastic conidiogenesis of Botrytis cinerea; (d) conidiophores arising from aggregated hyphae, the sporodochia of Fusarium sp., macro- and micoconidium (d1); (e) septated conidia of Bipolaris heveae (e1) and Cochliobolus verruculosus (e2). Scale bars: (a, b2) 10 mm; (b1, c, d1, e) 20 mm 226 Y. Gherbawy et al. Table 11.5 Sequences retrieved from GenBank for the reconstruction of phylogenetic trees (Fig. 11.4) Strain GenBank accession nos. ACT TEF 18S rDNA 28S rDNA Absidia caeruleaNT AJ287133 AF157226 AF113405 AF113443 Absidia glauca AJ287135 X54730 AF157118 AF157172 Blakeslea trispora AJ287143 AF157235 AF157124 AF157178 Chaetocladium brefeldii AJ287144 AF157236 AF157125 AF157179 Chaetocladium jonesii AJ287145 AF157237 AF157126 AF157180 Choanephora cucurbitarum AJ287147 AF157239 AF157127 AF157181 Circinella umbellata AJ287148 AF157240 AF157128 AF157182 Dichotomocladium elegans AJ287153 AF157245 AF157131 AF157185 AJ287158 AF157250 AF157135 AF157189 Fennellomyces linderiT Gilbertella persicaria AJ287159 AF157251 AF157136 AF157190 Halteromyces radiatusT AJ287161 AF157253 AF157138 AF157192 Lichtheimia corymbifera AJ287134 AF157227 AF113407 AF113445 Lichtheimia hyalosporaT AJ287132 AF157225 AF157117 AF157171 Lichtheimia ramosa EU826377 EU826382 EU826361 EU826370 Mortierella alpina EU736236 EU736263 EU736290 EU736317 Mortierella multidivaricata AJ287168 AF157260 AF157144 AF157198 Mortierella verticillata AJ287170 AF157262 AF157145 AF157199 Mucor circinelloides AJ287173 AF157264 AF113427 AF113467 Mucor mucedo AJ287176 AF157267 X89434 AF113470 AJ287177 AF157268 AF113430 AF113471 Mucor racemosusT Mycotypha africanaIT AJ287180 AF157271 AF157147 AF157201 Mycotypha microspora AJ287181 AF157272 AF157148 AF157202 Parasitella parasitica AJ287182 AF157273 AF157149 AF157203 AJ287183 AF157274 AF157150 AF157204 Phascolomyces articulosusT Phycomyces blakesleeanus AJ287184 AF157275 AF157151 AF157205 Protomycocladus faisalabadensis AJ287189 AF157280 AF157156 AF157210 Radiomyces spectabilis AJ287190 AF157281 AF157157 AF157211 Rhizomucor miehei AJ287191 AF157282 AF113432 AF113473 Rhizomucor pusillus AJ287192 AF157283 AF113433 AF113474 Rhizopus arrhizus AJ287198 AF157289 AF113440 AF113481 Rhizopus stolonifer AJ287199 AF157290 AF113441 AF113482 AJ287200 AF157291 AF113442 AF113483 Saksenaea vasiformisT Spinellus fusiger AJ287201 AF157292 AF157159 AF157213 Syncephalastrum monosporum AJ287203 AF157294 AF157161 AF157215 Syncephalastrum racemosum AJ287204 AF157295 X89437 AF113484 Thamnostylum piriforme AJ287207 AF157298 AF157164 AF157218 AJ287208 AF157299 AF157165 AF157219 Thermomucor indicae-seudaticaeT Umbelopsis isabellina AJ287209 AF157300 AF157166 AF157220 Umbelopsis nana AJ287210 AF157301 AF157167 AF157221 Umbelopsis ramanniana AJ287166 AF157258 X89435 AF113463 AJ287212 AF157303 AF157169 AF157223 Zychaea mexicanaT T-type strain; IT-isotype strain; NT-neotype strain were retrieved from GenBank as references: AB113022 and AB113023 (R. stolonifer var. stolonifer CBS150.83 and CBS609.82), DQ119009 (R. microsporus var. chinensis, CBS631.82 type), DQ119011 (R. microsporus var. oligosporus CBS339. 62), DQ119014 (R. microsporus var. rhizopodiformis IP676.72), DQ119010 11 Molecular Barcoding of Microscopic Fungi with Emphasis Mortierella multidivaricata Mortierella verticillata Mortierella alpina Umbelopsis isabellina 100 Umbelopsis ramanniana 100 Umbelopsis nana 100 Phycomyces blakesleeanus Spinellus fusiger 100 Thermomucor indicae-seudaticae 100 100 Rhizomucor miehei 100 Rhizomucor pusillus Dichotomocladium elegans 100 100 100 Lichtheimia corymbifera 100 Lichtheimia hyalospora 100 Lichtheimia ramosa Protomycocladus faisalabadensis 100 Syncephalastrum monosporum 99 Syncephalastrum racemosum 100 100 Fennellomyces linderi 100 Thamnostylum piriforme 97 Zychaea mexicana 100 100 Circinella umbellata Phascolomyces articulosus Saksenaea vasiformis Radiomyces spectabilis 100 Halteromyces radiatus 100 Absidia caerulea 100 100 Absidia glauca 100 100 Rhizopus arrhizus Rhizopus stolonifer 100 100 Mycotypha africana Mycotypha microspora Gilbertella persicaria 100 100 Blakeslea trispora 100 Choanephora cucurbitarum 100 Mucor mucedo Mucor circinelloides 100 Chaetocladium brefeldii 100 100 Chaetocladium jonesii 73 Mucor racemosus 0.1 substitutions/ site 100 Parasitella parasitica 227 Mortierellales Umbelopsidaceae Phycomycetaceae Mucoraceae Syncephalastraceae Lichtheimiaceae Hoffmann et al., 2009 Syncephalastraceae Mucoraceae Syncephalastraceae Radiomycetaceae Absidiaceae (von Arx, 1982) Hoffmann et al., 2007 Mucoraceae Mycotyphaceae Choanephoraceae Mucoraceae Fig. 11.4 Bayesian inferred phylogram based on aligned nucleotide sequences encoding actin, translation elongation factor 1alpha, small and large subunit ribosomal RNA from 38 mucoralean fungi with 3 species of the Mortierellales as outgroup (see Table 11.5). Family affiliations are according to Kirk et al. (2008), Hoffmann et al. (2007, 2009b). Branch support values (posterior probabilities) are given and branch support values equal or greater than 75% in all analyses are indicated as bold branches (R. microsporus var. mircrosporus IP1124.75), DQ641325 (R. caespitosus CBS427.87), DQ641324 (R. homothallicus CBS336.62 type), DQ119031 (R. arrhizus var. arrhizus CBS112.07 type), AY213687 (R. schipperae CBS138.95 type), AB113016 (R. sexualis CBS336.39 type), DQ118991 (M. circinelloides f. circinelloides CBS195.68 neotype), AJ271061 (M. circinelloides f. lusitanicus CBS277.49), and DQ118984 (Lichtheimia corymbifera CBS120805). Furthermore, appropriate 228 Y. Gherbawy et al. sequences were retrieved from the genome projects of M. circinelloides f. lusitanicus CBS277.49 (scaffold 3 and 12, as of March-29-2009; http://www.jgi.doe.gov) and Rh. arrhizus var. arrhizus (supercontig 3.6 2078935-2079563 and 2042075-2042703, as of March-29-2009; http://www.broad.mit.edu/annotation/genome/rhizopus_ oryzae). The phylogenetic analysis of the ITS sequences was based on the Bayesian inference and is shown in Fig. 11.5. 11.2.7 Random Amplified DNA Polymorphisms Genomic DNA from several isolates of R. stolonifer, R. arrhizus, and M. circinelloides (Table 11.6) were amplified with the primers V6 (Lopandic et al. 1996) and M13 (O’Donnell et al. 1999). The random amplified DNA polymorphism (RAPD) profiles obtained were manually transferred into a binary data matrix (0 and 1 for absence and presence of RAPD bands) and subjected to distance based UPGMA analysis using PAUP* v4.0b10 (Swofford 1998). The combined matrices of both RAPD analyses consist of 49 characters in the case of Mucor and 68 characters for Rhizopus. RAPD analyses and corresponding trees are displayed in Figs. 11.6 and 11.7. 11.2.8 Sequence Similarity Matrices for Rhizopus and Mucor sp. Sequence similarity matrices were generated from the alignments, which were also used for the phylogenetic reconstructions. But it turned out that missing characters need to be omitted, and thus the aligned DNA matrix needs to be shortened to equal ends (Tables 11.7 and 11.8). 11.3 Results and Discussion 11.3.1 Diversity and Coarse Scale Identification of Fungal Species Isolated From Saudi Arabian Soil and Fruits A total of 120 different fungi were isolated, divided into Basidiomycota with 8 isolates, Ascomycota with 75 isolates out of 20 different genera with at least 26 species, and “Zygomycota” with 37 isolates belonging to 8 different genera and 13 species (Tables 11.1–11.4). Each phylum is characterised by specific features allowing a relative rough and easy assignment of its species. Often no more than a light microscope is necessary for the first rough identification. More or less 11 Molecular Barcoding of Microscopic Fungi with Emphasis 229 Lichtheimia corymbiferaCBS120805 Lichtheimia corymbiferaFSU6250 Rhizopus schipperae CBS138.95 T R. caespitosus CBS427.87 99 100 R. homothallicus CBS336.62 T 100 R. microsporus var.microsporus IP1124.75 100 100 R. microsporus var. chinensis CBS631.82 T R. microsporus var.oligosporus CBS339.62 R. microsporus var. rhizopodiformis IP676.72 R. arrhizus var. arrhizus CBS112.07 T 100 100 R. arrhizus var. arrhizus genome sc3.6 2078935-2079563 R. arrhizus var. arrhizus genome sc3.6 2042075-2042703 R. arrhizus TUR5 100 R. arrhizus TUR9 R. sexualis CBS336.39 T R. stolonifer CBS150.83 100 R. stolonifer TUR1 100 R. stolonifer TUR6 R. stolonifer TUR7 R. stolonifer CBS609.82 100 Mucor circinelloides f. circinelloides CBS 195.68 NT 100 M. circinelloides f. circinelloides TUM1 AM933548 M. circinelloides f. circinelloides FSU6518 M. circinelloides f. circinelloides FSU6251 93 M. circinelloides f. circinelloides FSU6252 M. circinelloides f. griseo-cyanus FSU6257 M. circinelloides f. griseo-cyanus FSU6259 88 M. circinelloides f. griseo-cyanus FSU9637 M. circinelloides f. griseo-cyanus TUM3 100 M. circinelloides f. griseo-cyanus TUM10 M. circinelloides f. griseo-cyanus TUM15 M. circinelloides f. griseo-cyanus TUM6 M. circinelloides f. griseo-cyanus TUM12 M. circinelloides f. griseo-cyanus TUM7 M. circinelloides f. griseo-cyanus TUM14 M. circinelloides genomescaffold 3 100 M. circinelloides f. lusitanicus CSB277.49 0.1 substitutions/ site M. circinelloides genome scaffold 12 Fig. 11.5 Bayesian inferred phylogram of aligned ITS sequences from different species and subspecies of Mucor circinelloides and Rhizopus sp. Branch support values are Posterior Probabilities. Strain numbers are given. Type or neotype strains are indicated by “T” or “NT” 230 Y. Gherbawy et al. Table 11.6 Rhizopus sp. and Mucor circinelloides strains isolated from different substrates Species R. arrhizus R. arrhizus R. stolonifer R. stolonifer R. stolonifer M. circinelloides M. circinelloides M. circinelloides M. circinelloides M. circinelloides Isolation code TUR5 TUR9-10 TUR1-4 TUR6 TUR7-8 TUM1-2 TUM3-4 TUM5-10 TUM11-14 TUM15 Substrate Date Soil Apricot Plum Grape Apricot Date Plum Grape Soil R. stolonifer 1 (apricot) R. stolonifer 2 (apricot) R. stolonifer 3 (apricot) R. stolonifer 4 (apricot) R. stolonifer 7 (grape) R. stolonifer 6 (grape) R. stolonifer 8 (grape) R. stolonifer 5 (date) R. stolonifer 9 (soil) R. stolonifer 10 (soil) RAPD - M13 RAPD - V6 Fig. 11.6 RAPD analyses of several isolates of Rhizopus sp. from Saudi Arabia. The profiles of two different primers, M13 and V6, were combined and subjected to distance based UPGMA analysis. The combined matrix consists of 68 characters regularly septated mycelia are typical for the hyphal growth of Asco- and Basidiomycota (Fig. 11.1a) in contrast to the nearly unseptated (but if septae are present, than irregularly septated) mycelium of the “Zygomycota”. Eponymous for each phylum are the sexual structures of reproduction, namely basidium, ascus (Fig. 11.1c), and zygospore (Fig. 11.1d). For microscopic fungi with only a limited number of distinctive, phylogenetic relevant morphological characters, the asexually developed mitospore- and their associated structures, possess great diagnostic importance. Although the Basidiomycetes isolated within this study did not form any distinctive features, e.g. anamorphic structures or even fruiting bodies in culture, the occurrence of regularly septated mycelium with clamps was diagnostic for dikaryotic mycelia of the Basidiomycetes (Fig. 11.1a, b). Because of the lack of suitable morphological parameters, a continuing morphological identification was not possible for any of the basidiomycetous isolates. A more detailed identification was attempted using nucleotide sequence data of the nuclear ITS region. Three 11 Molecular Barcoding of Microscopic Fungi with Emphasis 231 M. circinelloides 15 (soil) M. circinelloides 1 (apricot) M. circinelloides 7 (plum) M. circinelloides 8 (plum) M. circinelloides 3 (date) M. circinelloides 4 (date) M. circinelloides 9 (plum) M. circinelloides 10 (plum) M. circinelloides 11 (grape) M. circinelloides 12 (grape) M. circinelloides 13 (grape) M. circinelloides 14 (grape) M. circinelloides 2 (apricot) M. circinelloides 5 (plum) M. circinelloides 6 (plum) RAPD - M13 RAPD - V6 Fig. 11.7 RAPD analyses of several isolates of Mucor circinelloides from Saudi Arabia. The profiles of two different primers, M13 and V6, were combined and subjected to distance based UPGMA analysis. The combined matrix consists of 49 characters isolates could be assigned to order and family level, and with good BLAST results to the species level, namely Phlebia radiata and Merulius tremellosus. (Table 11.1). For both species, the BLAST search was unequivocal for the genus level, but because of the lack of reliable reference sequences, the species level delimitation remains still somewhat uncertain. The other five isolates could not be assigned by ITS to order or below-order level. As a result, neither ITS sequences nor morphological traits are sensitive enough to differentiate between the anamorphic stages of basidiomycetes, if obvious distinctive features of the hyphae are missing. For that purpose the development of alternative gene markers are mandatory. Asco- and Zygomycetes differ in their type of mitospore formation. An exogenous sporulation is typical for Ascomycetes but atypical for Zygomycetes. The latter ones produce their mitospores endogenously within sporangia, sporangiola, and merosporangia. The sporangia in the order Mucorales are more or less globose (to subglobose) with a distinctive columella of varying size and shape (Fig. 11.2). The columella is synapomorphic for the Mucorales (Voigt et al. 2009). One of the most prominent exogenous spore disposals is the deliberation from special conidiogenous cells, the phialides in Ascomycetes. Phialides arise either directly from the conidiogenous cell (Fig. 11.3a) or from metulae (Fig. 11.3b). Conidial disposal can also take part from polyblastic conidiogenous cells, which arrange terminally on tree-like branches (Fig. 11.3c). Other ascomycetes bear their conidiophores on aggregated hyphae, the sporodochia (Fig. 11.3d). The ascomycetous mitospores appear in various shapes, types, and cellular integrities. While the 232 Table 11.7 Sequence identity matrix of ITS sequences from different isolates of Rhizopus species R. a. R. R. R. R. R. R. a. genome caespitosus homothallicus microsporus shipperae sexualis type R. caespitosus ID R. homothallicus 76 ID R. microsporus 73 77 ID R. shipperae 63 61 59 ID R. sexualis 45 44 48 44 ID R. arrhizus type 69 70 67 64 47 ID 69 71 67 64 47 99 ID R. arrhizus genome 46 47 49 45 67 48 48 R. stolonifer CBS150.83 R. stolonifer 43 45 47 42 60 46 46 isolate 1 46 47 49 45 66 48 48 R. stolonifer isolates 6&7 67 68 65 62 48 97 98 R. arrhizus isolates 5&9 R. st. R. st. CBS150.83 isolate 1 R. st. R. a. isolates 6&7 isolates 5&9 ID 84 ID 99 84 ID 48 46 48 ID Y. Gherbawy et al. 11 M.c. f. g-c. FSU6257 M.c. f. g-c. FSU6259 M.c. f. g-c. FSU9637 ID 99 ID 99 99 Molecular Barcoding of Microscopic Fungi with Emphasis Table 11.8 Sequence identity matrix of ITS sequences from different isolates of Mucor circinelloides M.c. f. c. M. c. f. g-c. M. c. f. g-c. M. c. f. l. M. c. f. c. M. c. f. c. M.c f. c.. genome neotype isolate 1 FSU6251 FSU6252 isolate 3 iso.6..//..15 ID M. c. f. lusitanicus genome ID M. c. f. circinelloides 95 NT 100 ID M. c. f. circinelloides 96 isolate 1 97 98 ID M. c. f. circinelloides 94 FSU6251 99 99 97 ID M. c. f. circinelloides 95 FSU6252 M. c. f. griseo92 95 95 93 94 ID cyanus isolate 3 M .c. f. g-c. iso. 91 94 94 92 93 98 ID 6,7,10,12,14,15 96 98 98 96 97 95 94 M. c. f. griseocyanus FSU6257 M. c. f. griseo96 99 99 96 98 96 94 cyanus FSU6259 96 98 98 96 96 95 95 M. c. f. griseocyanus FSU9637 ID 233 234 Y. Gherbawy et al. mitospores of the ascomycetes can be uni- or multicellular, septated or unseptated (Figs. 11.3d1 and e), the mitospores of the zygomycetes are always unicellular and unseptated. On the basis of the anatomy of 75 ascomycete isolates, 74 (99%) could be identified to genus level, but only 43 (57%) could be identified to species level. From 64 species the ITS region, from 27 species the beta-tubulin gene, and from 8 species the calmodulin gene were sequenced. Those 99 sequences were subjected to BLAST searches. In all cases an identification of the genus was possible. An assignment to species level was possible for 56% of the ITS sequences, 82% for the nucleotide sequences encoding beta-tubulin, and 100% for those encoding calmodulin. Consequently, the application of the ITS as barcode marker is ambivalent. While ITS works nicely for molecular barcoding of the ascomycetous genera Trichoderma, Hypocrea, or Trichophyton (Druzhinina et al. 2005; Summerbell et al. 2007), it is not variable enough to distinguish between species in the genera Penicillium (Skouboe et al. 1996, 1999) and Fusarium (possessing non-orthologous copies of ITS2, O’Donnell and Cigelnik 1997; O’Donnell et al. 1998). In Aspergillus, the ITS is quite useful to separate different sections from each other but also not useful to distinguish between species in Aspergillus (Balajee et al. 2007). Obviously, the species delimitations are narrower than in other genera. Alternative barcode markers were evaluated and successfully applied. These are partial sequences of the gene encoding translation elongation factor 1alpha (tef) for Fusarium (Geiser et al. 2004) and the genes encoding beta-tubulin or calmodulin for the most studied ascomycetous genera (Geiser et al. 2007; Hong et al. 2006; Samson et al. 2004). For Zygomycetes, the situation is somewhat different. Here the morphological markers are discriminative enough to gain a proper above and below species-level identification for all isolates. On the other hand, the ITS sequences facilitated a proper identification of genera, but not of species. Only 80% zygomycetes could be reliably identified down to the species level. The missing fifth (20%) of misapplied classifications occurred because of (1) missing reference material (e.g. for M. varians and M. circinelloides f. griseo-cyanus), or (2) morphologically problematic reference and type strains (FSU9654; M. hiemalis and Rhizomucor variabilis are morphologically quite similar; see discussion in Hoffmann et al. 2009a). But by carefully judging these possible problems, ITS possesses a high sensitivity and is a very powerful tool for a rough classification of Zygomycetes, but tends to be more precise in combination with phenotypic criteria. Thus, a single method of identification should always be supplemented with an additional support by different unlinked nucleotide sequence barcode markers in combination with morphological and physiological markers. Nevertheless, an advantage of ITS over others is that the primers are universally applicable and the sequences are usually diverse enough to distinguish between taxa down to the species level. A major drawback of the ITS lies in its repetitive nature and in the resulting escape of single copies from concerted evolution leading to different ITS sequence types in Zygomycetes (Schwarz et al. 2006). However, a critical vision of all paralogs of the ITS increases the quantity and the quality of the discriminating signal for zygomycete identification (Alastruey-Izquierdo et al. 2010). 11 Molecular Barcoding of Microscopic Fungi with Emphasis 235 In summary, apart from the choice of the right molecular barcode marker, the most serious problems in the molecular identification are still caused by a lack in the availability of suitable reference sequences (Hoffmann et al. 2009a) and their annotation in the public data bases of the International Nucleotide Sequence Database Collaboration (Nilsson et al. 2006). If sequence material is thought to be the primary information source for identification of isolates as intended in the various international barcode projects, an all-encompassing database of perfectly described, annotated voucher specimens is indispensable. Such an effort for an organised molecular identification of fungi was initiated by the creation of Mycobank at www.mycobank.org, an initiative of the Centraalbureau voor Schimmelcultures Utrecht in the Netherlands (Crous et al. 2004). Mycobank provides onward links to DNA databases and nomenclatural novelties accessible in Mycobank, IndexFungorum, GBIF, and other international biodiversity initiatives for the realisation of a species bank that eventually links all databases of life. 11.3.2 Diversity and Morphological Identification of the Mucoralean Genera Rhizopus and Mucor Within the order Mucorales (Mucoromycotina) two prominent genera exist, namely Rhizopus and Mucor. Whereas Rhizopus comprises closely related taxa (BS 100% in Fig. 11.4; Voigt et al. 1999; Voigt and Wöstemeyer 2001), the genus Mucor forms a highly polyphyletic group (Fig. 11.4; O’Donnell et al. 2001; Hoffmann et al. 2009a; Voigt et al. 2009). Both genera harbour species of ubiquitous soil fungi mainly living as saprobes. But under certain conditions some species could be responsible not only for agricultural and food spoilage but also for mucormycoses in humans and animals (Michailides and Spotts 1990; Ogawa et al. 1992; Ribes et al. 2000; Voigt et al. 1999). The genus Rhizopus EHRENBERG with its type R. stolonifer is characterised by apophysate sporangia, stolons, and rhizoids. The sporangiophores arise mostly opposite the rhizoids (Fig. 11.2b3; Ehrenberg 1820; Schipper 1984; Schipper and Stalpers 1984). Traditional classification was mainly based on morphology (Zycha et al. 1969). In a revision, the genus was divided into three groups: the R. arrhizus (formerly R. oryzae) group, the R. microsporus group, and the R. stolonifer group (Schipper 1984; Schipper and Stalpers (1984). R. arrhizus represents a single species group; whereas the group of R. microsporus harbours its varieties, R. m. var. microsporus, R. m. var. rhizopodiformis, R. m. var. oligosporus, R. m. var. chinensis, and also R. homothallicus. The R. stolonifer group contained initially two species, R. stolonifer and R. sexualis comprising the varieties R. s. var. stolonifer, R. s. var. lyococcus, and R. s. var. sexualis and R. s. var. americanus, respectively (Schipper and Stalpers 1984). But later the varieties were elevated at the rank of species, e.g. R. americanus (Zheng et al. 2000) and R. lyococcus (Liou et al. 2007). On the basis of 28S rDNA (LSU, D1/D2 region), the members 236 Y. Gherbawy et al. of the R. lyococcus group can be clearly distinguished from the R. stolonifer group. This is concordant with the appearance of recurved sporangiophores exclusively found in the R. lyococcus group (Liou et al. 2007). In a monograph on the genus Rhizopus morphological traits, growth temperature, mating behaviour, and molecular systematics were considered as the main descriptive characters (Zheng et al. 2007). The authors accepted ten species and seven varieties, but no groups. The Dictionary of the Fungi (Kirk et al. 2008) and the database of IndexFungorum (www.indexfungorum.org; as of 31th August 2009) recognise nine to ten species, respectively and eight varieties in the genus Rhizopus. Out of 155 entries in IndexFungorum, six entries were re-classified in four other genera, twelve entries were synonymous for R. stolonifer, and more than 70 entries were synonyms for R. arrhizus var. arrhizus. From the latter species, the clinical isolate RA99-880 (FGSC9543; NRRL43880) was sequenced and published by the Broad Institute of the Harvard University and the Massachusetts Institute of Technology (MIT). This species was chosen as one important representative agent of mucormycoses in order to gain knowledge on genes and their impact on pathobiology (www.broad.mit. edu). R. arrhizus differs from R. stolonifer by smaller sized and darker coloured sporangia, by the ornamentation of the zygospores, and by its high thermotolerance allowing colonisation of warm-blooded animals including humans (Schipper 1984). Within the present study, several isolates from fruits collected in Saudi Arabia could be identified as members of the groups R. stolonifer and R. arrhizus. Identification was performed on the basis of micromorphology and DNA polymorphisms generated by nucleotide sequences and DNA fingerprints (Figs. 11.5 and 11.6). On the other hand, the genus Mucor FRESENIUS with its type M. mucedo is characterised by the formation of non-apophysate sporangia (Fig. 11.2a) and the lack of stolons and rhizoids (Fresenius 1850; Schipper 1975). As the genus Mucor is highly polyphyletic ( O’Donnell et al. 2001; Voigt and Wöstemeyer 2001), the number of accepted species varies between 50 and 75 (Kirk et al. 2008; www.indexfungorum.org; as of 31th August 2009). Out of currently 700 entries in IndexFungorum, 116 entries were re-classified in forty other genera and fifteen were synonyms for the type species M. mucedo. The genome of M. circinelloides f. lusitanicus CBS 277.49 was sequenced by the Joint Genome Institute (www.jgi.doe.gov). M. circinelloides is currently divided into four formae with a total of 25 synonyms. The formae circinelloides, lusitanicus, janssenii, and griseo-cyanus differ mainly in their colony colour, sporangiospores, and appearance of the columellae (Schipper 1976). Differentiation between these formae still remains largely on morphological traits. Nucleotide sequence comparisons of the different formae are reported Sect. 11.3.5 of this chapter. M. circinelloides is not closely related to any other species of Mucor. Its closest relatives are M. mucedo and M. racemosus as well as the facultative parasites Parasitella and Chaetocladium (100% BS, Fig. 11.4). 11 Molecular Barcoding of Microscopic Fungi with Emphasis 237 11.3.3 Phylogenetic Relevance of Morphological Markers While the common morphological characteristics of both genera, Rhizopus and Mucor, turned out to be very similar, only those of Rhizopus bear phylogenetic relevance, which allows a secure assignment to the genus. The combination of the criteria “stolons and rhizoids, apophysate sporangia, sporangiophores arising opposite the rhizoids” define a well-supported distinct evolutionary group, whereas the appearance of “non-apophysate sporangia and the lack of stolons and rhizoids” is not enough to describe the genus Mucor, because it matches also with the main characteristics of many other genera polyphyletic in Mucor (Fig. 11.4). As outlined in Sect. 11.3.1. the construction of an online platform linking all databases harbouring nucleotide sequences, as well as distinctive morphological and growth physiological criteria of the type strains from the core Mucoraceae, will help to revise the genus delimitations towards a natural system reflected in a monophylogenetic concept of the taxa. 11.3.4 Fine Scale Identification of Rhizopus Isolates Based on Combined Morphological and Molecular Characters Although various species of Rhizopus are morphologically very similar, there are some phylogenetic applicable characters allowing affiliations to different species complexes and species. Especially the complexity of rhizoids, length of sporangiophores, size of sporangia, appearance of zygospores, temperature range for growth, and preferred substrates differentiate between the three species complexes (as outlined in detail by Schipper and Stalpers 1984). Thus, the species of Rhizopus isolated within this study were unequivocally identified by the application of the morphological traits mentioned before. The species designations were successfully supported by analyses of DNA sequence (Tables 11.3–11.4). 11.3.5 Synergistic Application of DNA-Polymorphism and DNA-Sequence Generating Tools Based on RAPD analyses, the isolates of R. arrhizus differ largely from those of R. stolonifer (Fig. 11.6). Two isolates of R. arrhizus from soil (samples 9 and 10) were identical in their banding patterns but differed largely from isolate 5, which was collected from a date fruit (Fig. 11.6). Nevertheless, the analysis of the ITS1 and 2 revealed no differences between the isolates from soil and date (Fig. 11.5, Table 11.7). In a sequence identity matrix, these isolates of R. arrhizus show highest similarities (97–98%) to the type of R. arrhizus CBS112.07 and to the 238 Y. Gherbawy et al. strain RA99-880 used in the genome project (Table 11.7). R. arrhizus is the sister group to R. microsporus complex with the species R. microsporus, R. caespitosus, and R. homothallicus. R. schipperae appears basal to R. arrhizus / R. microsporus supported by 99% PP (Fig. 11.5). According to sequence similarities and the original description, R. schipperae belongs clearly to the R. microsporus complex (Table 11.7; Weitzman et al. 1996). Sequence similarities within the R. microsporus complex range between 73–77% (except R. schipperae with ca. 60%) and are around 62–71% to the group of R. arrhizus (Table 11.7). The R. microsporus group and the R. arrhizus group are together the sister group to the R. stolonifer group including R. sexualis (100% PP, Fig. 11.5; Schipper 1984; Schipper and Stalpers 1984). RAPD analyses of different isolates from Saudi Arabian fruits show clearly two distinctive groups, each with more or less similar banding patterns, but on an average, the isolates collected from apricots show fewer bands than isolates collected from grapes (Fig. 11.6). ITS-sequence similarities within the species R. stolonifer are 84% between isolates from apricots and grapes, 84% between the apricot isolate and strain CBS150.83, and 99% between the grape isolate and CBS150.83 (Table 11.7). All isolates show similarities of 60–67% to R. sexualis. The differences between R. sexualis and R. stolonifer are similarly compared to those of R. microsporus versus R. arrhizus, thus indicating well-described and supported taxonomic group affiliations. Species of the genus Rhizopus seem to develop a certain host-specificity (Fig. 11.6). This is not the case in M. circinelloides (Fig. 11.7). Although there are obvious correlations between the analyses of ITS sequences and RAPDs for Rhizopus isolates, there is no further correlation between these two analyses of the different isolates of Mucor circinelloides. On the basis of the ITS sequences, the isolates 6, 7, 10, 12, 14, and 15 are identical (data not shown, 100%) with closest similarity to isolate 3 (98%, Table 11.8). But an unequivocal assignment to different formae could not be achieved. Although the two formae isolated from Saudi Arabia were differentiated on the basis of the morphological criteria as M. circinelloides f. circinelloides and M. circinelloides f. griseo-cyanus, there is no reference sequence available from the latter forma. Both formae could be separated by their different size of the sporangiospores and the shape of their columellae. On the basis of ITS sequence overall similarities, only the forma lusitanicus is clearly distinguishable from both other formae with 4–9% sequence dissimilarities (Table 11.8). But the sequence of the neotype from M. circinelloides f. circinelloides CBS195.68 shows 98–99% sequence identity to M. circinelloides f. griseocyanus FSU6257, FSU6259, and FSU9637. This similarity is higher than that of 97% with the isolate FSU6251, which was identified as M. circinelloides f. circinelloides (Table 11.8). Isolates of M. circinelloides f. griseo-cyanus show sequence dissimilarities between themselves ranging from 1–6% and sequence dissimilarities from 1–8% to the isolates of M. circinelloides f. circinelloides. This high deviations in sequence similarities is below forma-level and does therefore not facilitate a sufficient differentiation between the two formae. But evaluating the sequence alignment, few nucleotide positions occur that may help to tell the formae apart (Fig. 11.8). Within the 550 basepair long ITS1-5.8S rDNA-ITS2 alignment, 15 11 Molecular Barcoding of Microscopic Fungi with Emphasis 239 Fig. 11.8 ITS sequence alignment of several isolates of Mucor circinelloides. Alignment length is 550 base pairs. Alignment is cut to the positions with variable sites. Dots indicate invariant sites of the sequences. Nucleotide sequence positions which are typical for the different formae are highlighted in black. Formae are abbreviated as follows: “l” – lusitanicus; “gc”– griseo-cyanus; “c”– circinelloides nucleotide positions are obvious for M. circinelloides f. lusitanicus and clearly separate this forma from M. circinelloides f. circinelloides and M. circinelloides f. griseo-cyanus. These positions are position 5 (C for M. circinelloides f. lusitanicus instead of T in the other formae = C/T), position 26 (C/T), position 44 (C/T), position 54 (T/C), position 171 (G/C), position 362 (gap/A), position 379 (T/A), position 380 (G/T), position 399 (G/T), position 400 (gap/A), position 401 (gap/T), position 404 (A/T), position 449 (gap/A), position 545 (C/T), and position 449 (gap/TC) (Fig. 11.8). For differentiation of the other two formae griseo-cyanus and circinelloides, only few distinctive nucleotide positions exist. At position 86, a cytosine is characteristic for M. circinelloides f. griseo-cyanus instead of a thymidine in the other formae. An insert near the end of the alignment seems typically for M. circinelloides f. griseo-cyanus, but is not present in all isolates of this formae. M. circinelloides f. circinelloides show 3 characteristic nucleotide positions that differentiate this form from the other forms. At position 100 a guanosine is localised 240 Y. Gherbawy et al. as a substitute for an adenosine. At position 367 is an inserted thymidine and at position 382 a thymidine instead of a cytosine. Such specific nucleotide positions within aligned sequences serve as markers, which differentiate between species and subspecies when the whole sequence is not able to separate properly. Unlike Rhizopus and its fruit-dependent ITS variances, there are no such differences within the isolates and formae of M. circinelloides and their substrate collection. Primer V6 shows no correlations between the origin of the isolate, e.g., with similar patterns for isolates 3 & 4 (from date) and 9 &10 (from plum) (Fig. 11.7). Dissimilar patterns were also observed, e.g., the isolated pairs 7 & 8, 9 & 10, and 5 & 6, which all originate from plums (Fig. 11.7). Even there are more bands for the RAPD primer M13; there is also no correlation between origin and below-species level designation possible. But primer M13 could differentiate between e.g. 3 & 4 (date) and 9 & 10 (plum). The similar RAPD patterns for both primers between the pairs 3 & 4, 5 & 6, 9 & 10, 11 & 12, and 13 & 14 suggest that these species originate from a common clonal line and do presumably belong to M. circinelloides f. griseocyanus (Fig. 11.7).Thus, the species of Rhizopus are more host-specific than the isolates of Mucor circinelloides. Consequently, it can be argued that host specificity may take part at the species level rather than at the below-species level. More detailed information about the molecular identification of food- and fruitborne Rhizopus and Mucor strains are published elsewhere (Gherbawy and Hussien 2009). 11.3.6 Evaluation of Potential Barcoding Methods and the Impact of Extended Species Recognition Species of the investigated genera Rhizopus and Mucor harbour important agents of post harvest diseases and mucormycoses. A fast, accurate, cost-saving method to unequivocally assign an unknown isolate to a species would be of great interest. Although there are only few species, which cause mucormycoses (mainly Rhizopus arrhizus and Mucor circinelloides), there exist different sensitivities against antifungals (Ribes et al. 2000; Dannaoui et al. 2003; Alastruey-Izquierdo et al. 2009). Thus, profound studies on the geographic occurrence and diversity, genetic variability, and epidemiological significance of important fungal species are required. DNA barcodes like ITS, IGS, 28S rDNA sequences and RAPDs already proved to be useful markers to differentiate species and varieties for well-supported and wellstudied fungal groups (this study; Liou et al. 2007; Liu et al. 2008; Hoffmann et al. 2009a). However, molecular barcoding based on short standardised DNA regions is neither a tool for phylogenetic reconstructions nor for taxonomical purposes. Molecular barcoding solely provides means for the linkage of a sample specimen to already existing taxonomical, systematic, and phylogenetic information. But a non-negligible advantage of molecular barcoding is its ability for aiding ecologists to gain insights into species diversity and identity of environmental samples, a goal in well-timed pest control (Valentini et al. 2009). 11 Molecular Barcoding of Microscopic Fungi with Emphasis 241 11.3.7 The Pressing Need for Reliable Species Identification The present systematics of Zygomycetes suggests a more generalised lifestyle than that in derived fungi. Host specificity is manifested already on species level, whereas many derived fungi show broad host ranges on species level and small host ranges on sub-species levels. For instance, the ascomycetous genus Fusarium is well-studied because of its agricultural importance as causative agent of the head blight of wheat as well as crown and stem rots causing immense yield losses and mycotoxin envenomations in humans and animals after the consumption of infected grains. Epidemiological studies revealed a very broad host range with different formae speciales of Fusarium spp. based on their specificities to certain host plants (Michielse and Rep 2009). Probably because of the underestimation of the plant and animal damaging potential of zygomycetous fungi, the number of currently described species is most likely undervalued. With a denser monitoring of their diversity and their epidemiology the identification of a broader spectrum of species also defined by substrate specificity can be predicted. In recent years several studies were published, reducing existing species to the status of “synonyms” resulting in the aforementioned high number of synonyms (e.g. Hoffmann et al. 2007, 2009a, b; Alastruey-Izquierdo et al. 2010; Zheng et al. 2007). For example, more than 70 synonyms were described for R. arrhizus (Zheng et al. 2007; IndexFungorum). Few physiological and chemotaxonomical characters not commonly used for the identification of Zygomycetes, such as differences in the utilisation of carbon and nitrogen sources, the formation of specific compounds (cell wall sugars, fatty acids etc.), or the quantity of sporulation, were not accepted by the scientific community leading to a considerable reduction of the number of species by Zycha (1935) and others. On the other hand, the ability to produce organic acids was recently considered in species recognition, separating the R. arrhizus group once again into R. arrhizus var. arrhizus and R. arrhizus var. delemar (Abe et al. 2003; Zheng et al. 2007). Substrate specificities combined with established and accepted morphological and phylogenetic approaches could be a useful tool to differentiate and restore already described species to accomplish a natural system of species, which does also consider evolutionary distances in order to gain an increase in the objectivity of setting species limitations. The resulting refinement in the definition of the phylogenetic species concept (Taylor et al. 2000) allows a more reliable and precise identification of a fungal specimen and is essential for comparative verifications of industrial, environmental, or pathogenic strains (Ogawa et al. 1992; Hachmeister and Fung 1993; Voigt et al. 1999; Hageskal et al. 2006). 11.4 Conclusions For the identification and the classification of specimens, a conscientious plan of the approaches with careful judgement of the obtained data is mandatory and will lead to sustainable results. The combination of different scientific expertise will solve the puzzle of diverse characters. 242 Y. Gherbawy et al. With the concatenation of traditional with molecular approaches and its advantage of universal applicability, the identification of fungi will rapidly develop and achieve a more precise delimitation of taxonomical groups than single criteria. But what could be the contribution of the DNA barcodes? The database of the Consortium for the Barcode of Life (http://www.barcoding.si.edu/) is also supposed to help non-taxonomists in the assignment of an unknown specimen to a welldescribed voucher specimen based on short barcode sequences, which are easy to obtain. These barcodes are thought of for the protection of the Earth’s biodiversity which is preceded by the elucidation of the Earth’s biodiversity. Molecular barcodes are essential where no other opportunities for identification are possible, e.g. missing discriminating data or under axenic conditions not cultivable specimens. The protection of endangered life forms, and the control of pests and their vectors are only few of the objectives examples supporting barcode research (CBOL, http:// barcoding.si.edu). But we should not forget that DNA barcodes were not designed for the identification of new species. DNA barcodes are often not sufficient enough for the exploration of new species and will probably fail (Aliabadian et al. 2009). Nevertheless, with the ability to identify fungal species on the basis of sequence information, a trend can be detected in disregarding characters other than DNA sequences for the identification and the description of species. Trusting in a single or few non-descriptive characters is a highly alarming trend. For instance, single BLAST searches could be using inaccurate sequence data (Balajee et al. 2009; Bridge et al. 2003). Although there are constant changes in the terminology of fungal morphology and taxonomy, descriptive traditional characters are essential and must not be neglected. Furthermore, molecular, biochemical, ecological, and physiological parameters can efficiently reveal morphologically identical species, cryptic species. The exploration of cryptic species is accelerating in recent years, for example in the dikaryomycotean genera Trichoderma (Gams and Bissett 1998), Fusarium (Baayen et al. 2000; O’Donnell 2000; O’Donnell et al. 2000), Armillaria (Pegler 2000), Aspergillus (Geiser et al. 1998; Pringle et al. 2005), or Penicillium. With the identification and the development of characters which are independent from external influences, and therefore, less error-prone than other criteria, the DNA markers are often preferred over morphological data. Because of the occasionally large intra-specific variability, the lack of sufficient distinctive characters, or the dependence upon physiological parameters (Schipper 1973; Zycha et al. 1969), informative phenotypic data are often hard to obtain and to reproduce. But without reliable morphological data, an assignment of molecular data to a specific reference fungal specimen is not possible (Hoffmann et al. 2009a). Without reliable reference strains and reference sequences a molecular identification is impossible. With the ongoing trend to identifying fungal specimens using DNA barcodes, a broad-ranged and well-defined taxonomic database is crucial (Meyer and Paulay 2005) because barcodes are only useful to assign unidentified specimens to already known and described species. The same way, DNA markers can distinguish morphological cryptic species; molecular biologically cryptic species can be differentiated 11 Molecular Barcoding of Microscopic Fungi with Emphasis 243 by morphology. Therefore, the combination of phenotypic and molecular data is highly recommended. A gap located between the proposed total number of fungal species of at least 1.5 million (Hawksworth 1991; Kirk et al. 2001; Hawksworth 2001) and the number described so far, which ranges between 72,000 and 120,000 (Hawksworth and Rossman 1997; Hawksworth 2001), implies a large number of still undiscovered and undescribed species. Once the problems of DNA barcoding, resulting from insufficient sampling of voucher specimens, insufficient barcode markers (e.g. paraloges), or bad taxonomy of voucher specimens because of misidentification or because poly- or paraphyletic species are avoided, the large scaled molecular identification of fungi will be a powerful and rapid procedure for the assessment of the fungal organisms in the biosphere (Hebert et al. 2004a; Wiemers and Fiedler 2007; Funk and Omland 2003; Moritz and Cicero 2004). Authors’ Contributions and Competing Interests The authors declare that they have no competing interests. CK is a graduate student and performed identification by traditional analyses as well as sequencing of all fungi except the species related to the Rhizopus sp. – Mucor circinelloides part. YG and NG collected all Saudi Arabian fungal isolates and did all experiments related to the Rhizopus sp. – Mucor circinelloides part, including identification, sequencing, and RAPD analyses. YG provided inspiring ideas and contributed to the discussion of the results. KH performed the phylogenetic analyses, coordinated the work, and wrote the manuscript. Acknowledgements The authors wish to thank Martin Kirchmair (University of Innsbruck, Austria) who kindly provided the zygomycetes strains FSU6510-6527 and Roman Schwarz (Medizinisches Versorgungszentrum, Mönchengladbach, Germany) who kindly provided the zygomycetes strains FSU6529, FSU6530, FSU6250-6252, FSU6274. References Abe A, Sone T, Sujaya I-N, Saito K, Oda Y, Asano K, Tomita F (2003) rDNA ITS sequences of Rhizopus oryzae: its application to classification and identification of lactic acid producers. Biosci Biotechnol Biochem 67:1725–1731 Alastruey-Izquierdo A, Castelli MV, Cuesta I, Monzon A, Cuenca-Estrella M, Rodriguez-Tudela JL (2009) Activity of posaconazole and other antifungal agents against Mucorales strains identified by sequencing of internal transcribed spacers. Antimicrob Agents Chemother 53:1686–1689 Alastruey-Izquierdo A, Hoffmann K, Hoog de GS, Rodriguez-Tudela JL, Voigt K, Bibashi E, Walther G (2010) Species recognition and clinical relevance of the zygomycetous genus Lichtheimia (syn. Mycocladus, Absidia pp). in revision Aliabadian M, Kaboli M, Nijman V, Vences M (2009) Molecular identification of birds: performance of distance-based DNA barcoding in three genes to delimit parapatric species. PloS ONE 4:e4119 Almyroudis N, Sutton D, Fothergill A, Rinaldi M, Kusne S (2007) In vitro susceptibilities of 217 clinical isolates of Zygomycetes to conventional and new antifungal agents. Antimicrob Agents Chemother 51:2587–2590 244 Y. Gherbawy et al. Alvarez E, Sutton DA, Cano J, Fothergill AW, Stchigel A, Rinaldi MG, Guarro J (2009) Spectrum of zygomycete species identified from clinically significant specimens in the United States. J Clin Microbiol doi:101128/JCM00036-09 Arnot DE, Roper C, Bayoumi RA (1993) Digital codes from hypervariable tandemly repeated DNA sequences in the Plasmodium falciparum circumsporozoite gene can genetically barcode isolates. Mol Biochem Parasitol 61:15–24 Baayen RP, O’Donnell K, Bonants PJM, Cigelnik E, Kroon LPNM, Roebroeck EJA, Waalwijk C (2000) Gene genealogies and AFLP analyses in the Fusarium oxysporum complex identify monophyletic and nonmonophyletic formae speciales causing wilt and rot disease. Phytopathology 90:891–900 Bal AM (2006) Importance of identication of zygomycetes in the era of newer antifungal agents. Transpl Infect Dis 8:122–123 Balajee SA, Houbraken J, Verweij PE, Hong S-B, Yaghuchi T, Varga J, Samson RA (2007) Aspergillus species identification in the clinical setting. Stud Mycol 59:39–46 Balajee SA, Borman AM, Brandt ME, Cano J, Cuenca-Estrella M, Dannaoui E, Guarro J, Haase G, Kibbler CC, Meyer W, O’Donnell K, Petti CA, Rodriguez-Tudela JL, Sutton D, Velegraki A, Wickes BL (2009) Sequence-based identification of Aspergillus, Fusarium and Mucorales in the clinical mycology laboratory: where are we and where should we go from here? J Clin Microbiol 47:877–884 Benjamin RK (1979) Zygomycetes and their spores. In: Kendrick B (ed) The whole fungus the sexual-asexual synthesis. National Museum of Natural Science. National Museums of Canada and The Kananaskis Foundation Ottawa, Canada, pp 573–616 Berbee LM, Taylor JW (2001) Fungal molecular evolution: gene trees and geologic time In: McLaughlin DJ, McLaughlin EG, Lemke PA (eds) The Mycota, vol 7B. Systematics and evolution, Springer, Berlin, Germany, pp 229–245 Bradley RD, Baker RJ (2001) A test of the genetic species concept: cytochrome b sequences and mammals. J Mammal 82:960–973 Bridge PD, Roberts PJ, Spooner BM, Panchal G (2003) On the unreliability of published DNA sequences. New Phytol 160:43–48 Chase MW, Salamin N, Wilkinson M, Dunwell JM, Kesanakurthi RP, Haidar N, Svolainen V (2005) Land plants and DNA barcodes: short-term and long-term goals. Philos Trans R Soc London B 360:1889–1895 Crous PW, Gams W, Stalpers JA, Robert V, Stegehuis G (2004) Mycobank: an online initiative to launch mycology into the 21st century. Stud Mycol 50:19–22 Cuenca-Estrella M, Gomez-Lopez A, Mellado E, Buitrago MJ, Monzon A, Rodriguez-Tudela JL (2006) Head-to-head comparison of the activities of currently available antifungal agents against 3, 378 Spanish clinical isolates of yeasts and filamentous fungi. Antimicrob Agents Chemother 50:917–921 Dannaoui E, Meletiadis J, Mouton JW, Meis JF, Verweij PE, The Eurofung Network (2003) In vitro susceptibilities of Zygomycetes to conventional and new antifungals. J Antimicrob Chemother 51:45–52 de Hoog GS, Guarro J, Gené J, Figueras MJ (2000) Atlas of clinical fungi. 2nd edn. Centraalbureau voor Schimmelcultures/University Rovira i Virgili, Utrecht, The Netherlands,/Reus, Spain Diwakar A, Dewan RK, Chowdhary A, Randhawa HS, Khanna G, Gaur SN (2007) Zygomycosis – a case report and overview of the disease in India. Mycoses 50:247–254 Druzhinina IS, Kopchinskiy AG, Komon M, Bissett J, Szakacs G, Kubicek CP (2005) An oligonucleotide barcode for species identification in Trichoderma and Hypocrea. Fungal Genet Biol 42:813–828 Ehrenberg CG (1820) De mycetogenesi as Acad C.L.C.N.C Praesidem Epistola Nova Acta Academiae Caesareae Leopoldino-Carolinae Germanicae. Naturae Curiosorum 10:159–222 Einax E, Voigt K (2003) Oligonucleotide primers for the universal amplification of beta-tubulin genes facilitate phylogenetic analyses in the regnum Fungi. Org Divers Evol 3:185–194 Ellis JJ, Hesseltine CW (1965) The genus Absidia: globose-spored species. Mycologia 57:222–235 11 Molecular Barcoding of Microscopic Fungi with Emphasis 245 Ellis JJ, Hesseltine CW (1966) Species of Absidia with ovoid sporangiospores II. Sabouraudia 5:59–77 Felsenstein J (1985) Confidence limits of phylogenies: an approach using the bootstrap. Evolution 39:783–791 Fenchel T, Finlay BJ (2006) The diversity of microbes: resurgence of the phenotype. Philos Trans R Soc B 361:1965–1973 Fresenius G (1850) Mucor mucedo. Beiträge zur Mykologie 1:7 Funk DJ, Omland KE (2003) Species-level paraphyly and polyphyly: frequency, causes, and consequences, with insights from animal mitochondrial DNA. Annu Rev Ecol Evol Syst 34:397–423 Gams W, Bissett J (1998) Morphology and identification of Trichoderma In: Kubicek CP, Harman GE (eds) Trichoderma and Gliocladium. Basic Biology, Taxonomy and Genetics. Taylor and -Francis, London, UK, pp 3–34 Geiser DM, Pitt JI, Taylor JW (1998) Cryptic speciation and recombination in the aflatoxinproducing fungus Aspergillus flavus. Proc Natl Acad Sci USA 95:388–393 Geiser DM, Jimenez-Gasco MD, Kang SC, Makalowska I, Veeraraghavan N, Ward TJ, Zhang N, Kuldau GA, O’Donnell K (2004) FUSARIUM–ID v 10: a DNA sequence database for identifying Fusarium. Eur J Plant Pathol 110:473–479 Geiser DM, Klich MA, Frisvad JC, Peterson SW, Varga J, Samson RA (2007) The current status of species recognition and identification in Aspergillus. Stud Mycol 59:1–10 Gherbawy Y, Hussien N (2009) Molecular characterization of Mucor circinelloides and Rhizopus stolonifer strains isolated from some Saudi fruits. Foodborne Pathog Dis 7(2) doi: 10.1089/ fpd.2009.0359 Glass NL, Donaldson GC (1995) Development of primer sets designed for use with the PCR to amplify conserved genes from filamentous ascomycetes. Appl Environ Microbiol 61:1323–1330 Greenberg RN, Scott LJ, Vaughn HH, Ribes JA (2004) Zygomycosis (mucormycosis): emerging clinical importance and new treatments. Curr Opin Infect Dis 17:517–525 Hachmeister KA, Fung DY (1993) Tempeh: a mold-modified indigenous fermented food made from soybeans and/or ceral grains. Crit Rev Microbiol 19:137–188 Hageskal G, Knutsen AK, Gaustad P, deHoog GS, Skaar I (2006) Diversity and significance of mold species in Norwegian drinking water. Appl Environ Microbiol 72:7586–7593 Hawksworth DL (1991) The fungal dimension of biodiversity: magnitude, significance, and conservation. Mycol Res 95:641–655 Hawksworth DL (2001) The magnitude of fungal diversity: the 15 million species estimate revisited. Mycol Res 105:1422–1432 Hawksworth DL, Rossman AY (1997) Where are all the undescribed fungi? Phytopathology 87:888–891 Hebert PDN, Cywinska A, Ball SL, deWaard JR (2003) Biological identification through DNA barcodes. Proc R Soc Lond B 270:313–322 Hebert PDN, Stoeckle MA, Zemlak TS, Francis CM (2004a) Identification of birds through DNA barcodes. PLoS Biol 2:1657–1663 Hebert PDN, Penton EH, Burns JM, Janzen DH, Hallwachs W (2004b) Ten species in one, DNA barcoding reveals cryptic species in the neotropical skipper butterfly Astraptes fulgerator. Proc Natl Acad Sci USA 101:14812–14817 Helbig AJ, Seibold I (1999) Molecular phylogeny of Palearctic–African Acrocephalus and Hippolais warblers (Aves: Sylviidae). Mol Phylogenet Evol 11:246–260 Hesseltine CW, Ellis JJ (1961) Notes on Mucorales, especially Absidia. Mycologia 53:406–426 Hesseltine CW, Ellis JJ (1964) The genus Absidia: Gongronella and cylindrical-spored species of Absidia. Mycologia 56:568–601 Hesseltine CW, Ellis JJ (1966) Species of Absidia with ovoid sporangiospores I. Mycologia 58:761–785 Hibbett DS, Binder M, Bischoff JF, Blackwell M, Cannon PF, Eriksson OE, Huhndorf S, James T, Kirk PM, Lücking R, Lumbs HT, Lutzoni F, Matheny PB, McLaughlin DJ, Powell MJ, Redhead S, Schoch CL, Spatafora JW, Stalpers JA, Vilgalys R, Aime MC, Aptroot A, 246 Y. Gherbawy et al. Bauer R, Begerow D, Benny GL, Castlebury LA, Crous PW, Dai YC, Gams W, Geiser DM, Griffith GW, Gueidan C, Hawksworth DL, Hestmark G, Hosaka K, Humber RA, Hyde KD, Ironside JE, Kõljalg U, Kurtzman CP, Larsson KH, Lichtwardt R, Longcore J, Miadlikowska J, Miller A, Moncalvo JM, Mozley-Standridge S, Oberwinkler F, Parmasto E, Reeb V, Rogers JD, Roux C, Ryvarden L, Sampaio JP, Schüssler A, Sugiyama J, Thorn RG, Tibell L, Untereiner WA, Walker C, Wang Z, Weir A, Weiss M, White MM, Winka K, Yao YJ, Zhang N (2007) A higherlevel phylogenetic classification of the Fungi. Mycol Res 111:509–547 Higgins DG, Sharp PM (1988) CLUSTAL: a package for performing multiple sequence alignment on a microcomputer. Gene 73:237–244 Higgins DG, Sharp PM (1989) Fast and sensitive multiple sequence alignments on a microcomputer. Comp Appl Biosciences 5:151–153 Hoffmann K, Discher S, Voigt K (2007) Revision of the genus Absidia (Mucorales, Zygomycetes): based on physiological, phylogenetic and morphological characters: thermotolerant Absidia spp form a coherent group, the Mycocladiaceae fam nov. Mycol Res 111:1169–1183 Hoffmann K, Telle S, Walther G, Eckart M, Kirchmair M, Prillinger H, Prazenica A, Newcombe G, Dölz F, Papp T, Vágvölgyi C, deHoog S, Olsson L, Voigt K (2009a) Diversity, genotypic identification, ultrastructural and phylogenetic characterization of zygomycetes from different ecological habitats and climatic regions: limitations and utility of nuclear ribosomal DNA barcode markers. In: Gherbawy Y, Mach RL, Rai M (eds) Current advances in molecular mycology. Nova Science Publishers, Inc, New York, pp 263–312 Hoffmann K, Walther G, Voigt K (2009b) Mycocladus vs Lichtheimia: a correction (Lichtheimiaceae fam. nov., Mucorales, Mucoromycotina). Mycol Res 113:277–278 Hong SB, Go SJ, Shin HD, Frisvad JC, Samson RA (2005) Polyphasic taxonomy of Aspergillus fumigatus and related species. Mycologia 97:1316–1329 Hong SB, Cho HS, Shin HD, Frisvad JC, Samson RA (2006) Novel Neosartorya species isolated from soil in Korea. Int J Syst Evol Microbiol 56:477–486 Huelsenbeck JP, Ronquist F (2001) MRBAYES: bayesian inference of phylogeny. Bioinformatics 17:754–755 Iwen PC, Freifeld AG, Sigler L, Tarantolo SR (2005) Molecular identification of Rhizomucor pusillus as a cause of sinus-orbital zygomycosis in a patient with acute myelogenous leukaemia. J Clin Microbiol 43:5819–5821 Iwen PC, Sigler L, Noel RK, Freifeld AG (2007) Mucor circinelloides was identified by molecular methods as a cause of primary cutaneous zygomycosis. J Clin Microbiol 45:636–640 James TY, Kauff F, Schoch CL, Matheny PB, Hofstetter V, Cox C, Celio G, Gueidan C, Fraker E, Miadlikowska J, Lumbsch HT, Rauhut A, Reeb V, Arnold EA, Amtoft A, Stajich JE, Hosaka K, Sung G-H, Johnson D, O’Rourke B, Crockett M, Binder M, Curtis JM, Slot JC, Wang Z, Wilson AW, Schüßler A, Longcore JE, O’Donnell K, Mozley-Standridge S, Porter D, Letcher PM, Powell MJ, Taylor JW, White MM, Griffith GW, Davies DR, Humber RA, Morton J, Sugiyama J, Rossman AY, Rogers JD, Pfister DH, Hewitt D, Hansen K, Hambleton S, Shoemaker RA, Kohlmeyer J, Volkmann-Kohlmeyer B, Spotts RA, Serdani M, Crous PW, Hughes KW, Matsuura K, Langer E, Langer G, Untereiner WA, Lücking R, Büdel B, Geiser DM, Aptroot A, Diederich P, Schmitt I, Schultz M, Yahr R, Hibbett DS, Lutzoni F, McLaughlin D, Spatafora J, Vilgalys R (2006) Reconstructing the early evolution of the fungi using a six-gene phylogeny. Nature 443:818–822 James TY, Letcher PM, Longcore JE, Mozley-Strandridge SE, Porter D, Powell MJ, Griffith GW, Vilgalys R (2007) A molecular phylogeny of the flagellated fungi (Chytridiomycota) and description of a new phylum (Blastocladiomycota). Mycologia 98:860–871 Johnson NK, Cicero C (2004) New mitochondrial DNA data affirm the importance of Pleistocene speciation in North American birds. Evolution 58:1122–1130 Kirk PM, Cannon PF, David JC, Stalpers JA (2001) Ainsworth & Bisby´s Dictionary of the fungi, 9th edn. CABI Publishing, Wallingford, United Kingdom Kirk PM, Cannon PF, Minter DW, Stalpers JA (2008) Dictionary of the fungi, 10th edn. CABI Publishing, Wallingford, United Kingdom 11 Molecular Barcoding of Microscopic Fungi with Emphasis 247 Kress WJ, Erickson DL (2007) A two-locus global DNA barcode for land plants: the coding rbcL gene complements the non-coding trnHpsbA spacer region. PLoS ONE 2:e508 Kress WJ, Wurdack KJ, Zimmer EA, Weigt LA, Janzen DH (2005) Use of DNA barcodes to identify flowering plants. Proc Natl Acad Sci USA 102:8369–8374 Lemer S, Aurelle D, Vigliola L, Durand JD, Borsa P (2007) Cytochrome b barcoding, molecular systematics and geographic differentiation in rabbitfishes (Siganidae). C R Biol 330:86–94 Leslie JF, Summerell BA (2006) Fusarium laboratory manual, 1st edn. Ames, Iowa (ua), Blackwell Publishers, pp 1–388 Liu X, Huang H, Zheng R (2008) Delimitation of Rhizopus varieties based on IGS rDNA. Sydowia 60:93–112 Liou G, Chen S, Wei H, Lee F, Fu H, Yuan G, Stalpers JA (2007) Polyphasic approach to the taxonomy of the Rhizopus stolonifer group. Mycol Res 111:196–203 Lopandic K, Prillinger H, Molnar O, Gimenez -Jurado G (1996) Molecular characterisation and genotypic identification of Metschnikowia species. Syst Appl Microbiol 19:393–402 Metellus P, Laghmari M, Fuentes S, Eusebio A, Adetchessi T, Ranque S, Bouvier C, Dufour H, Grisoli F (2008) Successful treatment of a giant isolated cerebral mucormycotic (zygomycotic) abscess using endoscopic debridement: case report and therapeutic considerations. Surg Neurol 69:510–515 Meyer CP, Paulay G (2005) DNA barcoding: error rates based on comprehensive sampling. PLoS Biol 3:2229–2238 Michailides TJ, Spotts RA (1990) Postharvest diseases of pome and stone fruits caused by Mucor piriformis in the Pacific northwest and California. Plant Dis 74:537–543 Michielse CB, Rep M (2009) Pathogen profile update: Fusarium oxysporum. Mol Plant Pathol 10:311–324 Moritz C, Cicero C (2004) DNA barcoding: promise and pitfalls. PloS Biol 2:1529–1531 Newmaster SG, Fazekas AJ, Ragupathy S (2006) DNA barcoding in land plants: evaluation of rbcL in a multigene tiered approach. Can J Bot 84:335–341 Nilsson RH, Ryberg M, Kristiansson E, Abarenkov K, Larsson K-H, Kõljalg U (2006) Taxonomic reliability of DNA sequences in public sequence databases: a fungal perspective. PloS ONE 1:e59 Nucci M, Marr KA (2005) Emerging fungal diseases. Clin Infect Dis 41:521–526 O’Donnell K (2000) Molecular phylogeny of the Nectria haematococca – Fusarium solani species complex. Mycologia 92:919–938 O’Donnell K, Cigelnik E (1997) Two divergent intragenomic rDNA ITS2 types within a monophyletic lineage of the fungus Fusarium are nonorthologous. Mol Phyl Evol 7:103–116 O’Donnell K, Cigelnik E, Nirenberg HI (1998) Molecular systematics and phylogeography of the Gibberella fujikuroi species complex. Mycologia 90:465–493 O’Donnell K, Gherbawy Y, Schweigkofler W, Adler A, Prillinger H (1999) Comparative phylogenetic analysis of Fusarium oxysporum and related species from maize using DNA sequence and RAPD data. J Phytopathol 147:445–452 O’Donnell K, Kistler HC, Tacke BK, Casper HH (2000) Gene genealogies reveal global phylogeographic structure and reproductive isolation among lineages of Fusarium graminearum, the fungus causing wheat scab. Proc Natl Acad Sci USA 97:7905–7910 O’Donnell K, Lutzoni FM, Ward TJ, Benny GL (2001) Evolutionary relationships among mucoralean fungi (Zygomycota): evidence for family polyphyly on a large scale. Mycologia 93:286–296 Ogawa JM, Sonada RM, English H (1992) Postharvest diseases of tree fruit. In: Kumar J, Chaube HS, Singh US, Mukhopadhyay AN (eds) Plant diseases on international importance, vol III. Englewood Cliffs, Diseases of fruit crops Prentice Hall, pp 405–422 Park MH, Sim CJ, Baek J, Min GS (2007) Identification of genes suitable for DNA barcoding of morphologically indistinguishable korean halichondriidae sponges. Mol Cells 23:220–227 Pegler DN (2000) Taxonomy, nomenclature and description of Armillaria. In: Fox RTV (ed) Armillaria Root Rot: biology and control of honey fungus. Intercept, Andover, pp 81–93 248 Y. Gherbawy et al. Pringle A, Baker DM, Platt JL, Wares JP, Latge JP, Taylor JW (2005) Cryptic speciation in the cosmopolitan and clonal human pathogenic fungus Aspergillus fumigatus. Evolution 59:1886–1899 Raper KB, Fennell DI (1965) The genus Aspergillus. Williams and Wilkins Company, Baltimore, pp 1–686 Raper KB, Thom C (1949) Manual of the penicillia. Williams and Wilkins Company, Baltimore, pp 1–875 Ribes JA, Vanover-Sams CL, Baker DJ (2000) Zygomycetes in human disease. Clin Microbiol Rev 13:236–301 Roden MM, Theoklis TZ, Buchanan WL, Knudsen TA, Sarkisova TA, Schaufele RL, Sein M, Sein T, Chiou CC, Chu JH, Kontoyiannis DP, Walsh TJ (2005) Epidemiology and outcome of Zygomycosis: a review of 929 reported cases. Clin Infect Dis 41:634–653 Rogers TR (2008) Treatment of zygomycosis: current and new options. J Antimicrob Chemother 61:35–40 Rogers JS, Swofford DL (1998) A fast method for approximating maximum likelihood of phylogentic trees from nucleotide sequences. Syst Biol 47:77–89 Ronquist F, Huelsenbeck JP (2003) MRBAYES 3: bayesian phylogenetic inference under mixed models. Bioinformatics 19:1572–1574 Samson RA, Frisvad JC (eds) (2004) Penicillium subgenus Penicillium: new taxonomic schemes, mycotoxins and other extrolites. Stud Mycol 49:1–251 Samson RA, Seifert KA, Kuijpers AFA, Houbraken JAMP, Frisvad JC (2004) Phylogenetic analysis of Penicillium subgenus Penicillium using partial beta-tubulin sequences. Stud Mycol 49:175–200 Saunders GW (2005) Applying DNA barcoding to red macroalgae: a preliminary appraisal holds promise for future applications. Philos Trans R Soc Lond B 360:1879–1888 Schipper MAA (1973) A study on variability in Mucor hiemalis and related species. Stud Mycol (Baarn) 4:1–40 Schipper MAA (1975) Mucor mucedo, Mucor flavus and related species. Stud Mycol (Baarn) 10:1–33 Schipper MAA (1976) Mucor circinelloides, Mucor racemosus and related species. Stud Mycol (Baarn) 12:1–40 Schipper MAA (1984) A revision of the genus Rhizopus I The Rhizopus stolonifer-group and Rhizopus oryzae. Stud Mycol (Baarn) 25:1–19 Schipper MAA (1990) Notes on Mucorales – I Observations on Absidia. Persoonia 14:133–149 Schipper MAA, Stalpers JA (1984) A revision of the genus Rhizopus II The Rhizopus microsporus-group. Stud Mycol (Baarn) 25:20–34 Schüßler A, Schwarzott D, Walker C (2001) A new fungal phylum, the glomeromycota: phylogeny and evolution. Mycol Res 105:1413–1421 Schwarz P, Bretagne S, Gantier J-C, Garcı́a-Hermoso D, Lortholary O, Dromer F, Dannaoui E (2006) Molecular identification of Zygomycetes from culture and experimentally infected tissues. J Clin Microbiol 44:340–349 Seifert KA, Samson RA, deWaard JR, Houbraken J, Lévesque CA, Moncalvo J-M, Louis-Seize G, Hebert PDN (2007) Prospects for fungus identification using CO1 DNA barcodes, with Penicillium as a test case. PNAS 104:3901–3906 Singh J, Rimek D, Kappe R (2005) In vitro susceptibility of 15 strains of Zygomycetes to nine antifungal agents as determined by the NCCLS M38-A microdilution method. Mycoses 48:246–250 Skouboe P, Boysen M, Pedersen LH, Frisvad JC, Rossen L (1996) Identification of penicillium species using the internal transcribed spacer (ITS) regions. In: Rossen L, Rubio V, Dawson MT, Frisvad JC (eds) Fungal identification techniques. European Commission, Brussels, Belgium, pp 160–164 Skouboe P, Frisvad JC, Taylor JW, Lauritsen D, Boysen M, Rossen L (1999) Phylogenetic analysis of nucleotide sequences from the ITS region of terverticillate Penicillium species. Mycol Res 103:873–881 11 Molecular Barcoding of Microscopic Fungi with Emphasis 249 Sugiyama J (1998) Relatedness, phylogeny, and evolution of the fungi. Mycoscience 39: 487–511 Summerbell RC, Moore MK, Starink–Willemse M, Van Iperen A (2007) ITS barcodes for Trichophyton tonsurans and T equinum. Med Mycol 45:193–200 Swofford DL (1998) PAUP*: phylogenetic analysis using parsimony (*and other methods) version 4. Sinauer Associates, Sunderland, MA Tavares ES, Baker AJ (2008) Single mitochondrial gene barcodes reliably identify sister-species in diverse clades of birds. BMC Evol Biol 8:81 Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS, Fisher MC (2000) Phylogenetic species recognition and species concepts in fungi. Fungal Genet Biol 31:21–32 Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG (1997) The Clustal X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucl Acids Res 25:4876–4882 Trifilio SM, Bennett CL, Yarnold PR, McKoy JM, Parada J, Mehta J, Chamilos G, Palella F, Kennedy L, Mullane K, Tallman MS, Evens A, Scheetz MH, Blum W, Kontoyiannis DP (2007) Breakthrough zygomycosis after voriconazole administration among patients with hematologic malignancies who receive hematopoietic stem-cell transplants or intensive chemotherapy. Bone Marrow Transplant 39:425–429 Valentini A, Pompanon F, Taberlet P (2009) DNA barcoding for ecologists. Trends Ecol Evol 24:110–117 van de Peer Y, Baldauf SL, Doolittle WF, Meyer A (2000) An updated and comprehensive rRNA phylogeny of (crown) eukaryotes based on rate calibrated evolutionary distances. J Mol Evol 51:565–576 Vences M, Thomas M, Bonett RM, Vieites DR (2005) Deciphering amphibian diversity through DNA barcoding: chances and challenges. PhilOS Trans Soc Lond B 360:1859–1868 Vogelstein B, Gillespie D (1979) Preparative and analytical purification of DNA from agarose. Proc Natl Acad Sci USA 76:615–619 Voigt K, Wöstemeyer J (2001) Phylogeny and origin of 82 zygomycetes from all 54 genera of the Mucorales and Mortierellales based on combined analysis of actin and translation elongation factor EF-1alpha genes. Gene 270:113–120 Voigt K, Cigelnik E, O’Donnell K (1999) Phylogeny and PCR identification of clinical important Zygomycetes based on nuclear ribosomal-DNA sequence data. J Clin Microbiol 37:3957–3964 Voigt K, Hoffmann K, Einax E, Eckart M, Papp T, Vágvölgyi L, Olsson L (2009) Revision of the family structure of the Mucorales (Mucoromycotina, Zygomycetes) based on multigenegenealogies: Phylogenetic analyses suggest a bigeneric phycomycetaceae with spinellus as sister group to phycomyces. In: Gherbawy Y, Mach RL, Rai M (eds) Current advances in molecular mycology. Nova Science, New York, pp 263–312 Walsh TJ, Groll A, Hiemenz J, Fleming R, Roilides E, Anaissie E (2004) Infections due to emerging and uncommon medically important fungal pathogens. Clin Microbiol Infect 10:48–66 Weitzman I, McGough DA, Rinaldi MG, Della-Lata P (1996) Rhizopus schipperae, sp nov, a new agent of zygomycosis. Mycotaxon 59:217–225 Wheeler QD (2004) Taxonomic triage and the proverty of phylogeny. Philos Trans R Soc B 359:571–583 White TJ, Bruns T, Lee S, Taylor J (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ (eds) PCR protocols: a guide to methods and applications. Academic Press, San Diego, pp 315–322 White MM, James TY, O’Donnell K, Cafaro MJ, Tanabe Y, Sugiyama J (2006) Phylogeny of the Zygomycota based on nuclear ribosomal sequence data. Mycologia 98:872–884 Wiemers M, Fiedler K (2007) Does the DNA barcoding gap exist? – a case study in blue butterflies (Lepidoptera: Lycaenidae). Front Zool 4:8 Wollenweber HW, Reinking OA (1935) Die Fusarien, ihre Beschreibung. Schadwirkung und Bekämpfung Parey, Berlin, pp 1–355 250 Y. Gherbawy et al. Woo PC, Zhen H, Cai JJ, Yu J, Lau SK, Wang J, Teng JL, Wong SS, Tse RH, Chen R, Yang H, Lui B, Yuen KY (2003) The mitochondrial genome of the thermal dimorphic fungus Penicillium marneffei is more closely related to those of molds than yeasts. FEBS Lett 555:56–477 Yarlett N, Orpin CG, Munn EA, Yarlett NC, Greenwood CA (1986) Hydrogenosomes in the rumen fungus Neocallimastix patriciarum. Biochem J 15:729–739 Zheng RY, Chen GQ, Liu XY (2000) Rhizopus americanus stat. nov. Mycosystema 19:473–477 Zheng RY, Chen GQ, Huang H, Liu XY (2007) A monograph of Rhizopus. Sydowia 59:273–372 Zycha H (1935) Pilze II. Mucorineae. Krypt Fl Mark Brandenburg 6a, Leipzig Zycha H, Siepmann R, Linnemann G (1969) Mucorales. eine beschreibung aller gattungen und arten dieser pilzgruppe. J Cramer, Lehre Chapter 12 Advances in Detection and Identification of Wood Rotting Fungi in Timber and Standing Trees Giovanni Nicolotti, Paolo Gonthier, and Fabio Guglielmo Abstract Wood rotting fungi are reported as a major source of economic losses in both timber production and wood in service, and one of the main causes of tree wind throws and limb failures. Since the biology of these fungi is varied, their detection and identification are important for the application of appropriate management strategies and control measures. Following an overview of traditional and biochemical diagnostic techniques, whose usefulness is frequently limited either by their reliance on the sporadically emerging and rarely visible fruit bodies, or by the need of a preliminary isolation step, we discuss on DNA-based techniques that have been developed to detect and early identify wood rotting fungi in timber and in standing trees. 12.1 Introduction Wood rotting fungi, also named decay fungi or rots, are the primary biotic decomposers of the wood because of their ability to break down lignified cell walls (Blanchette 1991; Jasalavich et al. 2000). Except for few ascomycetes, all wood attacking fungi are basidiomycetes primarily belonging to Agaricales, Hymenochaetales, Polyporales, and Russulales (Kirk et al. 2001). Based on their enzymatic capabilities, the decay fungi are divided into brown rot fungi, which preferentially attack and rapidly depolymerize cellulose and hemicelluloses, and white rot fungi, which can progressively degrade both carbohydrates and lignin (Blanchette 1991; Worrall et al. 1997). The structural deterioration of wood, determined by both types of rots, is an essential process in carbon and nitrogen recycling of forest ecosystems, G. Nicolotti, P. Gonthier, and F. Guglielmo Di.Va.P.R.A., Department of Exploitation and Protection of the Agricultural and Forestry Resources, Plant Pathology, University of Torino, via L. da Vinci 44, I-10095, Grugliasco (TO), Italy e-mail: giovanni.nicolotti@unito.it Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_12, # Springer-Verlag Berlin Heidelberg 2010 251 252 G. Nicolotti et al. but can represent a serious threat leading to considerable problems in the urban context as well as in commercial forestry and plantations. In the urban environment, wood decay fungi can have a negative impact on tree stability (Lonsdale 1999). Furthermore, they are deemed to be responsible for most of damages of structural wood in buildings (Schmidt 2007). In forests and plantations for timber production and, to a lesser extent, in several fruit tree crops, root and butt rot diseases can cause severe economic losses (Bahnweg et al. 2002; Utomo and Niepold 2000). At any rate, timely detection and careful identification of wood rotting fungi are essential to define the most appropriate management strategies and control measures. Moreover, effective identification methods may be basic tools for studies focused on epidemiology, ecology, and biology of wood decay fungi. The late and occasional occurrence of visible rot symptoms and signs, such as the emergence of fruit bodies, can lead either to overlooking rot infections or to make identification of wood decay fungi unfeasible. This relevant issue has led to the development of several alternative techniques to efficiently detect and identify wood rotting fungi. Following a general overview of the problems and risks caused by rots in both standing trees and wood in service, as well as of the most hazardous fungi involved, this chapter is focused on methods for detection and identification of wood rotting fungi. Recently developed diagnostic techniques are discussed in terms of advances upon traditional methods, limits, and fields of application. Relevance is given to DNA-based techniques allowing for early identification of rots directly from wood. 12.2 Problems and Risks Related to Wood Rotting Fungi 12.2.1 Indoor Wood Rotting Fungi Indoor wood decay fungi, also named house rots, figure prominently amongst the economically most important wood inhabiting fungi (Schmidt 2007). Indeed, it has been estimated that the cost for repairing damages caused by rots to timber used in construction in 1977 in the UK amounted to £3 millions per week (Rayner and Boddy 1988). The rise of wood moisture content, due to general building defects or to the presence of permanent water vapor sources, can lead to environmental conditions suitable for fungal infections. Based on the extent of the damage and on the fungal species involved in the decay process, remedial treatments range from expensive refurbishment methods to a more practical disposal of moisture sources (Schmidt 2007). Not all indoor wood decay fungi are in fact problematic, and the knowledge of their physiological requirements is important to develop environmentally friendly control strategies (Högberg and Land 2004; Schmidt 2007). Identification of the house rots is therefore important for management purposes. Several studies have been focused on the abundance, economic significance, and biology of indoor wood inhabiting fungi (Bech-Andersen 1995; Gilbertson and 12 Advances in Detection and Identification of Wood Rotting Fungi 253 Ryvarden 1987; Huckfeldt and Schmidt 2006; Jellison et al. 2004). Brown rot basidiomycetes have been reported as the most common and hazardous indoor wood decay fungi. Among them, the dry rot fungi Serpula lacrymans (Wulfen:Fr.) and Meruliporia incrassata (Berk. & Curt.) are extremely destructive and among the least controllable fungi in Europe and the USA, respectively, mainly due to their ability to carry water and nutrients over long distances by means of strands (Jellison et al. 2004; Schmidt 2007). In Germany, S. lacrymans is clearly differentiated from other house rots as its presence requires considerably more rigorous and far-reaching control measures (Schmidt 2007). Other hazardous brown rot species include the cellar fungus Coniophora puteana (Schumach.) P. Karst., which proved to be able to cause decay even in wood with low moisture content (Huckfeldt and Schmidt 2006); the white polypores Antrodia spp. and Oligoporus placenta (Fr.) Gilb. & Ryvarden, mainly reported in the attic and upper floor; and the gill polypores Gloeophyllum spp., described as common destroyers of window and roof timber, with the ability to survive even at higher temperatures (Schmidt and Huckfeldt 2005). The Oak polypore, Donkioporia expansa (Desm.) Kotl. & Pouzar, has been reported as one of the few white rot basidiomycetes leading to relevant damages to indoor wood (Kleist and Seehann 1999). 12.2.2 Wood Rotting Fungi and Tree Stability An accurate inspection of trees in public areas can be essential for correct tree management plans aimed at preventing dangerous situations such as wind throws or limb failures. Although main stems and root systems of standing trees are structurally optimized to withstand several times the average of the mechanical forces to which they are subjected from their own weight and from loading by wind, rain, and snow (Mattheck and Breloer 1995), the structural biological deterioration of wood can increase the occurrence of mechanical failures. This may lead to severe damages to properties and to personal injuries (Lonsdale 1999). Moreover, excessive pruning and root lesions, to which urban trees are frequently exposed, may favor infections of wood rot fungi. A timely detection of hazardous trees is achieved through a careful visual inspection of decay signs and structural weakness (Mattheck and Breloer 1992). Although this method, complemented with instrumental analyses (Habermehl et al. 1999; Müller et al. 2001; Nicolotti et al. 2003; Tomikawa et al. 1990), can allow detecting wood rot even at an incipient stage, it rarely enables the diagnosis of the decay causative agent. Because the biology and ecology of different rots are varied, the knowledge of the wood decay fungi involved in each instance is important to predict, to some extent, the severity of the fungal infection and, thus, to reliably assess potential risks of failure (Lonsdale 1999). Early identification of the causative agent is then crucial for rapidly progressing decay fungi that can turn a sound tree into a hazard in a short period of time. Several comprehensive studies aimed at investigating and describing the most common rots in landscape and urban trees (Bernicchia 2005; Erkkilä and Niemelä 254 G. Nicolotti et al. 1986; Hickman and Perry 1997; Lonsdale 1999; Nicolotti et al. 2004a, b), as well as their invasiveness in different host species (Deflorio et al. 2008; Schwarze and Baum 2000; Schwarze et al. 2004; Swiecki et al. 2005; Terho et al. 2007), have allowed highlighting fungal taxa whose detection in a standing tree can be regarded as a hazard. Most of these taxa are white rot basidiomycetes: several species of Ganoderma and Inonotus have been reported as very active root and butt rot fungi potentially leading to extensive decay (Terho et al. 2007); Armillaria (Fr.:Fr.) Staude species are known to be dangerous root rot fungi occurring with high frequency even in urban environment (Guglielmo et al. 2008a); Phellinus spp. and Perenniporia fraxinea (Bull.) Ryvarden are deemed to be widespread butt and stem decay agents causing an intensive white rot (Bernicchia 2005; Nicolotti et al. 2004a, b; Swiecki et al. 2005). Although rarely found in hardwoods of urban environment, the brown rot Laetiporus sulphureus (Bull.) Murril is considered a hazardous root decay agent (Bernicchia 2005; Nicolotti et al. 2004a, b). Finally, the ascomycete Kretzschmaria deusta (Hoffm.) P.M.D. Martin has been recently proved to be strongly invasive in living stems of different deciduous hosts (Deflorio et al. 2008). 12.2.3 Rot Diseases and Timber Production Although root and butt rots represent one of the driving forces leading to spatial and temporal diversification of forests, rot diseases figure amongst the most prominent causes of cull in timber production, resulting in considerable economic losses worldwide (Delatour 1980; Hansen and Goheen 2000). Indeed, in conifer stands aimed at timber production, often characterized by intensive thinning and monoculture, rot frequency can amount up to 20% (Piri 1996). Since most of the rot fungi can display host preference or specificity, and peculiar spreading and infection biology, management practices and control measures to limit rot diseases in forests are strongly dependent on precise and accurate identifications of the disease agent. Several taxa within the Armillaria and Heterobasidion annosum (Fr.) Bref. species complexes are deemed to be responsible for most of the root and butt rot diseases of conifers and hardwoods in natural forest stands and plantations throughout the northern temperate regions of the world (Chase and Ullrich 1988; Kile et al. 1991). As an example, taxa included in the Heterobadision annosum species complex are reported to cause losses for more than €800 millions per year just in Europe (Woodward et al. 1998). The complex genus Armillaria encompasses about 40 biological species of varying geographic distributions, host ranges and virulence (Pegler 2000; Watling et al. 1991). Armillaria mellea (Vahl: Fries) Kummer and A. ostoyae (Romagnesi) Herink. have been reported as aggressive rot pathogens, whereas A. gallica Marxmuller and Romagnesi, A. cepistipes Velenovsky, A. borealis Marxmuller and Korhonen, and A. tabescens (Scopoli: Fries) Emel figure as secondary pathogens or weak parasites (Guillaumin et al. 1985; Kile et al. 1991; Wargo and Harrington 1991). Most of the Armillaria species are able to spread over long distances by means of rhizomorphs despite absence of root 12 Advances in Detection and Identification of Wood Rotting Fungi 255 contacts between adjacent trees. Before forest stand regeneration, careful removal of stumps or dying root systems is thus advisable to reduce fungal inoculum, especially when aggressive pathogenic Armillaria species are detected. Heterobasidion annosum sensu lato (s.l.) is a species complex including three European species, namely, H. abietinum Niemelä & Korhonen, H. annosum sensu stricto (s.s.), and H. parviporum Niemelä & Korhonen, and two North American intersterility groups (ISGs), H. annosum ISG P (Am-P) and H. annosum ISG S (Am-S). Each taxon within H. annosum s.l. is characterized by distinct host specialization: H. abietinum, H. parviporum, and Am-S have been reported mostly as butt rot agents on spruce or fir trees, whereas H. annosum s.s. and H. annosum Am-P are associated with root rot and mortality of pines (Korhonen and Stenlid 1998). Primary infection occurs by means of airborne basidiospores on fresh stumps or wounds, and secondary infection by vegetative spreading from stump to tree or from tree to tree through root contacts. Treating fresh stumps with chemicals or biotic competitors, as well as avoiding thinnings and clearcuttings in periods of abundant sporulation, are important preventive control measures against this pathogen (Gonthier et al. 2005; Möykkynen and Miina 2002; Nicolotti and Gonthier 2005). The identification of H. annosum species infecting a stand can be important for the selection of tree species to be used for reforestation (Hantula and Vainio 2003). Phellinus weirii (Murr.) Gilb. sensu lato and Inonotus tomentosus (Fr.) Teng are reported as aggressive root rot pathogens that can cause extensive wood losses and reduce productivity in conifer stands especially in North America (Germain et al. 2002; Hansen and Goheen 2000). Finally, although further investigations are needed, several species of Ganoderma, Phellinus, and Phlebia are responsible for harmful root and butt rot diseases affecting hardwood plantation for timber and biomass production in Asia (Lee 2000; Suhara et al. 2002; Utomo and Niepold 2000). 12.2.4 Root and Butt Rot Diseases and Fruit Tree Plantations Several root and basal stem rots have been reported as significant diseases in fruit tree plantations, and decay agents are capable of surviving in the soil for several years (Amenduni et al. 2001; Khairudin 1995). Curative measures are, in general, expensive and not always effective, owing to the fact that visible disease symptoms appear at a very late stage of infection (Utomo and Niepold 2000). Preventive strategies based on the use of pathogen-free propagating materials and planting in noninfested soils are the most appropriate control measures (Schena and Ippolito 2003). Detection and early identification tools are thus important to limit the spread of these diseases in fruit tree plantations. Rosellinia necatrix (Prill.) and A. mellea are dangerous root rot agents of fruit and forest trees with a widely distribution throughout temperate regions (Anselmi and Giorcelli 1990; Wargo and Harrington 1991). Basal stem rot caused by 256 G. Nicolotti et al. pathogenic Ganoderma spp. has been reported to severely reduce yearly harvest of oil palm crops in Asia (Utomo and Niepold 2000). 12.3 Traditional Techniques of Identification of Wood Rotting Fungi 12.3.1 Analysis of Fruit Bodies and Mycelial Strands Conventional identification methods of wood decay fungi mostly rely on visual analysis of fruit bodies. Dichotomous keys to species, available for several lignicolous basidiomycetes (Bernicchia 2005; Breitenbach and Kränzlin 1986; Hickman and Perry 1997; Hjortstam et al. 1978), are based on macromorphology of the basidioma and hymenophores, and on micromorphology of hyphal system, hymenial organs, and spores. Determination of fungal species through the use of these keys often requires a deep mycological background. Simplified keys for the identification of several wood rotting fungi in urban and landscape trees proved to be suitable for discrimination at the genus level (Intini et al. 2000; Lonsdale 1999; Strouts and Winter 1994). Because aggressiveness and ability to overcome host defenses can vary among different species within genera (Schwarze and Baum 2000), identification at a species level is often more useful for tree management purpose. Recently, field keys based on macroscopic observations of basidiomata, their longitudinal cross-sections, and color of spore print have been successfully validated and they include the most important and widespread European wood rot basidiomycetes in standing trees (Gonthier and Nicolotti 2007). Macroscopic and microscopic analysis of fruit bodies, as well as of mycelial strands, can be successfully used to identify house rots (Bravery et al. 2003; Huckfeldt and Schmidt 2006). Although this diagnostic method can be fast and reliable, it rarely allows for early identification of wood rotting fungi. Indeed, fruit bodies and/or mycelial strands of several wood decay fungi are rarely or sporadically visible and they usually emerge at advanced stages of the fungal infection (Palfreyman et al. 1991; Terho et al. 2007). This may represent a serious problem for several brown rots of wood in service, which can rapidly and drastically reduce wood strength at incipient stages, and for rapidly progressing root and butt rot agents of standing trees, which can represent a serious hazard in the urban environment. 12.3.2 Analysis of Pure Fungal Cultures Analysis of pure fungal cultures isolated from mycelium and/or decayed wood may be used when no fruit bodies are available. Keys based on growth rate, microscopic features, and enzymatic capabilities of mycelia have been published for the identification of several basidiomycetes at the species level (Lombard and Chamuris 12 Advances in Detection and Identification of Wood Rotting Fungi 257 1990; Nobles 1965; Stalpers 1978). Moreover, sexual or mating compatibility tests, by means of pairing the unknown isolate with known haploid testers, may be efficient diagnostic tools. A sexually compatible mating results in the production of a genetic stable heterokaryotic mycelium, visible both by changes in culture morphology or by microscopic observations for the presence of clamp connections at some or most septa (Harrington et al. 1989; Ullrich and Anderson 1978). Mating tests have been extensively used for the discrimination of species within Armillaria and Heterobasidion annosum species complexes (Anderson 1986; Korhonen 1978). Diagnostic methods based on pure culture analysis and tests are time-consuming and often unsuited to distinguish between closely related species (Schmidt 2007). Moreover, isolation of wood rotting fungi from environmental samples is often impractical, despite the use of selective media, due to the presence of fast growing fungal contaminants. 12.4 Biochemical Techniques 12.4.1 Protein-Based Techniques Identification of wood decay fungi by means of analysis of their proteins includes sodium dodecyl sulfate polyacrilamide gel electrophoresis (SDS-PAGE) and isozyme analysis. SDS-PAGE is based on the analysis of electrophoretic patterns of whole cell proteins, after denaturation by means of chemical treatments. This method has been used for the discrimination, at specific and subspecific levels, of several indoor wood decay fungi (Palfreyman et al. 1991; Schmidt and Moreth 1995). Analysis of isozymes, which are proteins with multiple forms but with similar or identical enzymatic properties, consists of an electrophoresis followed by treatment with dye-forming substrate for the target enzyme. The resulting isoenzymatic profiles can allow differentiation of closely related species. This technique has been extensively used for studies on inter and intraspecific variability within Armillaria and Heterobasidion annosum species complexes (Otrosina et al. 1992, 1993; Rizzo and Harrington 1993; Wahlstrom et al. 1991). Both the above-described techniques require the isolation of pure fungal cultures and a large amount of fungal tissues. Moreover, protein profiles can be highly affected by factors related to environment and the stage of fungal development. The procedure to develop systems yielding consistent results may then be work and time-consuming. 12.4.2 Immunological Techniques Immunological methods are based on the use of polyclonal antisera or monoclonal antibodies obtained to specifically recognize antigens, such as proteins or 258 G. Nicolotti et al. polysaccharides, typical of the species to be identified. Several methods, such as dot-blot immunoassay, enzyme-linked immunosorbent assay (ELISA), and electron microscopy immunolabeling have been developed for detection and/or quantification of the most serious indoor wood decay fungi (Clausen 1997; Jellison and Goodell 1988; Palfreyman et al. 2001). This technique proved to be useful to early detect rots directly from extracts of wood, without the need of any prior isolation and pure culturing step (Clausen and Kartal 2003). ELISA tests have been successfully developed for rapid detection from wood samples of A. mellea and A. ostoyae (Priestley et al. 1994) and pathogenic Ganoderma species on oil palm (Utomo and Niepold 2000). Although several immunological methods, such as ELISA, are promising techniques for screening a large number of samples, cross-reaction may occur with nontarget organisms (Schmidt 2007). Further, sensitivity of immunological assays can often be inhibited by wood extractives (Jellison and Goodell 1989). 12.5 DNA-Based Techniques The great potential of DNA-based techniques, over traditional and biochemical identification methods, relies on the chance to select diagnostic markers in coding as well as noncoding DNA regions of the nuclear and mitochondrial genome. Indeed, nucleotide sequence polymorphism of these regions can provide a large amount of diagnostic characters suitable for identification of fungi at different taxonomic levels independently of any factors related to environment and stage of fungal development. Although DNA hybridization techniques combined with the digestion of nuclear and mitochondrial DNA (nuc- and mt-DNA) by means of restriction endonucleases proved to be powerful for identification of Armillaria at a specific and subspecific rank (Jahnke et al. 1987; Anderson et al. 1989; Schulze et al. 1995), this method is too time-consuming and expensive for routine diagnosis of rots. Conversely, techniques based on polymerase chain reaction (PCR) are valuable alternative identification tools for specific, sensitive, and rapid routine diagnoses of wood decay fungi. As immunological methods, most PCR-based methods allow fungal identification directly from wood, without the need of any pure fungal culture isolation step. Following a brief overview of the protocols developed to efficiently extract DNA from wood, we describe the most important PCR-based methods used for identification of wood rotting fungi. 12.5.1 DNA Extraction from Wood While starting from fungal mycelium easy and rapid methods of hyphal suspension in sterile water followed by freezing and thawing steps may be suitable for PCR amplification of fungal ribosomal DNA (rDNA) (Garbelotto et al. 1996; Harrington 12 Advances in Detection and Identification of Wood Rotting Fungi 259 and Wingfield 1995); when wood samples have to be tested, more elaborate protocols are necessary for efficiently extracting fungal DNA. Wood extractives, namely polyphenols, tannins, resin acids, and polysaccharides, are known as potential PCR inhibitors (Bahnweg et al. 1998). Further, in decayed wood, DNA molecules can be subjected to partial degradation (Jasalavich et al. 2000). Efficient wood DNA extraction is thus crucial for both reliability and sensitivity of PCR-based methods. In a study focused on DNA isolation from recalcitrant materials, such as conifer tree roots and bark, Bahnweg et al. (1998) developed a highly efficient cetyltrimethylammonium bromide (CTAB) protocol consisting of an early extraction or precipitation of inhibiting components under conditions minimizing oxidation reactions. Since this method is work- and time-consuming, and requires several harmful reducing agents and organic solvents, it is unsuited to rapid routine diagnostics. Other protocols, based on serial extractions with either CTAB or SDS, and organic solvents (Jasalavich et al. 2000; Oh et al. 2003; Suhara et al. 2005; Vainio and Hantula 2000) efficiently provide amplifiable DNA from both incipient and advanced decayed wood but still result in long procedures. Rapid and organic solvent-free protocols consisting either of serial CTAB extractions (Råberg et al. 2005) or CTAB extraction followed by a final DNA purification with GENCLEAN Kit (Qbiogene, Carlsbad, CA, USA) (Allmér et al. 2006) or polyvinylpolypyrrolidone (PVPP) spin columns (Schena and Ippolito 2003) have proven to be useful in extracting fungal DNA from wood. Finally, in a comparative assay of different DNA extraction methods (Guglielmo 2005), a protocol entirely based on a kit developed to extract DNA from “stool” (Qiagen, Valencia, CA, USA) proved to be as efficient as the Bahnweg protocol for fungal DNA isolation from wood of different tree species. Since this method is rapid and no harmful reagent handling steps are necessary, it has been extensively and successfully used for the validation of PCR-based diagnostic methods on wood samples collected from different host tree species (Guglielmo et al. 2007, 2008b). Despite the cost, DNA extraction kits are very useful for reliable and rapid routine diagnostics. 12.5.2 RAPD-PCR Random amplified polymorphic DNA (RAPD) analysis consists of a PCR with an arbitrary and short oligonucleotide which can prime the amplification of DNA fragments when its complementary site occurs as reverted repeats in the genome (Williams et al. 1990). DNA polymorphism among different individuals is thus detected as the presence/absence of the amplicons that compose their RAPD profile. Schmidt and Moreth (1998) proposed RAPD markers that may be useful to distinguish S. lacrymans from other indoor decay fungi. Strain identification by means of RAPD analysis was paramount to prove the different origins of several C. puteana cultures that had been supposed to be strain Ebw15, an obligatory test fungus used to evaluate the efficacy of wood preservatives according to the European Standard EN113 (Göller and Rudolph 2003). RAPD analysis has been 260 G. Nicolotti et al. used to differentiate genotypes of Armillaria spp. in natural populations of different geographical areas, providing markers for the distinction of both a single clone (Smith et al. 1992) and a single species (Schulze et al. 1997). Finally, this method has been extensively used to study the genetic variability among and within ISGs of H. annosum (Garbelotto et al. 1993; Karjalainen 1996). Although RAPD analysis is fast and easy to perform and does not require prior information of the target DNA site, it can have limits of reproducibility. Moreover, this technique should not be applied on DNA extracted from environmental samples, i.e., wood, potentially holding several different microorganisms. 12.5.3 PCR-RFLP Restriction fragment length polymorphism (RFLP) analysis is based on the use of restriction endonucleases, enzymes which recognize specific nucleotide sequences (restriction sites) and consequently cut DNA at these or other points. Polymorphism in these sites can thus be used to distinguish between different individuals. As stated above, costly and elaborate Southern blotting and labeled probing techniques are needed to detect RFLP of total nuc- and/or mt-DNA. Conversely, RFLP applied to PCR-amplified DNA fragments can provide valuable diagnostic markers easily detectable through a simple agarose gel electrophoresis. Since nuc- and mtrDNA loci include both conserved and variable domains (Hong et al. 2002; White et al. 1990), they figure amongst the most popular DNA target sites for the development of most of the DNA-based fungal diagnostic techniques, including RFLP markers. Indeed, in the nuc-rDNA, internal transcribed spacers (ITSs) and nontranscribed intergenic spacers (IGSs) display high inter and intraspecific variability, whereas ribosomal genes, such as large and small rDNA (nuc-LrDNA and nuc-SrDNA), are more useful for identification at higher taxonomic levels (Bruns and Shefferson 2004; Guerin-Laguette et al. 2002). Multicopy arrangement and highly conserved priming sites, typical of both nuc- and mt-rDNA, allow DNA amplification from virtually all fungi, even if the starting sample is lacking in quantity or quality (Jasalavich et al. 2000; White et al. 1990). Amplified ribosomal DNA restriction analysis of ITS (ARDRA-ITS) has proven to be useful for the distinction of S. lacrymans from its closest relative Serpula hymantioides (Fr.) P. Karst. (Schmidt and Moreth 1999). Digestion of an IGS portion with a combination of restriction endonucleases allowed unambiguous identification of pure cultures of several Armillaria species from Europe (Sierra et al. 1999), North America (Harrington and Wingfield 1995; Sierra et al. 1999), and Japan (Matsushita and Suzuki 2005). RFLP of PCR-amplified ITS and IGS has been extensively used in several studies for the differentiation of H. annosum species and ISGs both in Europe and North America (Garbelotto et al. 1993; Gonthier et al. 2001; Kasuga and Mitchelson 2000; Kasuga et al. 1993). In a recent study, the development of RFLP of nuc-rDNA amplified fragments allowed the differentiation, within P. weirii complex, of P. weirii sensu stricto (s.s.) and 12 Advances in Detection and Identification of Wood Rotting Fungi 261 Phellinus sulphurascens Pilat, two species differing in pathogenicity but mostly undistinguishable by means of traditional methods (Lim et al. 2005). A wide collection of RFLP profiles based on a 1,800–1,900-bp-long region, including the ITS, allowed the differentiation among 48 out of 52 species of the European polypores analyzed (Fischer and Wagner 1999). A similar study, but restricted to RFLP analysis of ITS, has been performed to investigate wood inhabiting fungi in Picea abies of unmanaged forests (Johannesson and Stenlid 1999). PCR-RFLP has proved to be a valuable and reproducible method for fungal identification at different taxonomic levels, but as RAPD markers when applied to DNA extracted directly from wood it can lead to unreliable results. A cloning step before the digestion of the amplified rDNA (Kennedy and Clipson 2003) can overcome this limit but it makes the method longer and more complex. 12.5.4 T-RFLP Terminal restriction fragment length polymorphism (T-RFLP) is an automated version of PCR-RFLP. Fluorescently labeled primers are used in PCR and a highresolution capillary analyzer is used to investigate the digested fragments (Liu et al. 1997). In comparison to RFLP, only terminal digested fragments are detectable. This technique is a high-throughput fingerprinting method often used in studies of microbial communities (Edel-Hermann et al. 2004). Through T-RFLP on ITS, Coniophora puteana has been detected directly from artificially inoculated wood samples at early stages of colonization, when no hyphae were visible at the microscope (Råberg et al. 2005). T-RFLP, visual fruit body inspection, and analysis of pure fungal cultures have been simultaneously employed to examine the composition and the abundance of wood inhabiting fungi in woody debris of a Norway spruce stand (Allmér et al. 2006). Although the number of fungal species detected by T-RFLP was lower than that obtained with the other two methods, it allowed the detection of unculturable wood decay basidiomycetes (Lim et al. 2005). Although few applications of this method have been reported so far for diagnosis of wood rots, this technique is promising especially for rapid and simultaneous investigation of several fungi. 12.5.5 Taxon-Specific Priming PCR PCR with taxon-specific primers provides reliable tools for fungal diagnostics from both pure culture and environmental samples. The specificity of this method relies on the design of primers that anneal exclusively a complementary site unique for the taxon to be detected. If DNA of the target species is present, taxon-specific 262 G. Nicolotti et al. oligonucleotides prime the amplification of DNA fragment of a peculiar size. As an added benefit, the simultaneous application of taxon-specific primers in multiplex PCR reactions can increase the diagnostic capacity of PCR without compromising the specificity of the analysis (Elnifro et al. 2000). A simple agarose gel electrophoresis can allow the detection of taxon-specific amplicons. Species-specific oligonucleotides were designed on ITS to identify S. lacrymans, D. expansa, C. puteana, Antrodia vaillantii (DC.) Ryvarden, and Gloeophyllum sepiarium (Wulfen) P. Karst. (Moreth and Schmidt 2000; Schmidt and Moreth 2000). This technique can be useful to easily and rapidly detect, directly from wood samples, the economically most important basidiomycetes causing wood rot in European buildings (Table 12.1). PCRs with specific ITS primers have been developed to detect European Armillaria species by Schulze et al. (1997) and Lochman et al. (2004). Further, primers developed by Lochman et al. (2004) through a nested PCR method enable the detection of Armillaria spp. directly from environmental samples, such as soil. Attempts to design specific primers for each European Armillaria species are reported in the study conducted by Sicoli et al. (2003) (Table 12.2). A taxon-specific competitive priming (TSCP)-PCR, developed to identify in a single reaction the two north American H. annosum ISGs, Am-P and Am-S, allowed efficient typing of more than 500 fungal samples and to recover, first time in the field, a well-established hybrid between the two ISGs (Garbelotto et al. 1996). A similar approach has been used to study the abundance, potential dispersal range, and habitat of European H. annosum species in pure and mixed forests (Gonthier et al. 2001, 2003). While taxon-specific primers designed on mt-LrDNA with partially overlapping complementary sites allowed the distinction of H. parviporum and H. abietinum, the simultaneous use of universal fungal primers on the same region permitted the identification of the intronless H. annosum s.s. (Gonthier et al. 2003). For more practical purposes, PCR with primers designed on ITS allowed detecting both H. annosum s.s. and H. parviporum directly from increment cores of Norway Table 12.1 PCR assays, with reverse taxon-specific primers designed on ITS region, for the identification of indoor wood rotting fungi Primer pairs Primer sequence (50 -30 ) Specificity/Amplicon size References ITS1a tccgtaggtgaacctgcgg S. lacrymans/588 bp Moreth and Schmidt (2000) L aatgttgtcttgcgacaacg S. hymantioides/429 bp Moreth and Schmidt (2000) ITS1 tccgtaggtgaacctgcgg H tcccacaaccgaaacaaatc C. puteana/633 bp Moreth and Schmidt (2000) ITS1 tccgtaggtgaacctgcgg C agtagcaagtaaggcataga D. expansa/544 bp Moreth and Schmidt (2000) ITS1 tccgtaggtgaacctgcgg D tcgccaaaacgcttcacggt ITS1 tccgtaggtgaacctgcgg A. vaillantii/517 bp Moreth and Schmidt (2000) A caccgataagccgactcatt G. sepiarium/398 bp Moreth and Schmidt (2000) ITS1 tccgtaggtgaacctgcgg G gttaataaaaaccgggtgag a Universal forward primer designed by White et al. (1990) 12 Advances in Detection and Identification of Wood Rotting Fungi 263 Table 12.2 PCR assays for the identification of important root and butt rot agents of forest trees Primer pairs Primer sequence (50 -30 ) References Specificity/Amplicon DNA target site size Schulze et al. ARM-1 agggtatgtgcacgttcgac Armillaria spp./660 bpa ITS ARM-2 ggaaagctaagctcgcgcta (1997) AR-1 ctgacctgttaaagggtatgtgc Armillaria spp./690– ITS Lochman et al. AR-2 aagctgaatccttctacaaagtcaa (2004) 724 bpb IGS Sicoli et al. ATA1 ttgccttgaaccctgttataaggc A. tabescens/375– ATA2 tgccaaaatcgttgcacgccgc (2003) 381 bpb A. mellea/631 bp ITS Sicoli et al. AMEL3 ttgcttgcttacgagctaag ITS4c tcctccgcttattgatatgc (2003) A. mellea/364–387 bpb IGS Sicoli et al. AME1 aagaatcatgagatatcatcagt AME2 ttagaaaatccgccttagaaac (2003) gcatcgatgaagaacgcagc Armillaria spp./184 bp ITS Guglielmo ITS3c et al. (2007) Armi2R aaacccccataatccaatcc I. tomentosus/491 bp ITS Germain et al. It-ITS-209-f gctaaatccactcttaacac It-ITS-700-rc aggagccgaccacaaaagat (2002) PW164 gcttccatttttcttagg P. weirii s.s./495 bp ITS Lim et al. PW659 tcaaaagggcgtattaatg (2005) a The size is referred to the amplicon obtained from A. ostoyae DNA extracts b Depending on the isolates or species considered, the size of amplicon varies between the reported values c Universal primers designed by White et al. (1990) spruce (Bahnweg et al. 2002), whereas primers designed on cloned DNA fragments derived from random amplified microsatellites (RAMS) markers proved to be suitable to distinguish H. annosum s.s. and H. parviporum (Hantula and Vainio 2003). A summary of taxon-specific priming PCRs developed for the identification of H. annosum species is reported in Table 12.3. Taxon-specific primers were designed on ITS region for two other important root rot conifer pathogens, namely P. weirii s.s. and I. tomentosus (Germain et al. 2002; Lim et al. 2005) (Table 12.2), as well as for other significant hardwoods pathogens, such as Ganoderma species causing the basal stem rot of oil palm (Utomo et al. 2005); Phlebia brevispora Nakasone, involved in butt rot of Japanese cypress (Suhara et al. 2005); and Phellinus noxius, a destructive pathogen of several woody plants in Asia (Tsai et al. 2007). Three multiplex PCRs based on 11 taxon-specific primers designed on either nuc- or mt-rDNA regions allowed detecting several decay fungi, such as Armillaria spp., Ganoderma spp., Inonotus spp., K. deusta, Laetiporus spp., P. fraxinea, and Phellinus spp., reported as hazardous for tree stability in Europe and North America (Guglielmo et al. 2007; Nicolotti et al. 2009). Two further multiplex PCR reactions were developed to detect and identify at subgeneric levels the most hazardous wood decay fungi belonging to Ganoderma, Phellinus, and Inonotus (Guglielmo et al. 2008b). These methods were validated in the field and proved to be highly sensitive, allowing the detection of down to 10 1 pg of target DNA per 1 mg wood DNA extracts (Guglielmo et al. 2007). A summary of these multiplex PCRs, with related taxon-specific primers, is reported in Tables 12.4 and 12.5. 264 Table 12.3 TSCP-PCR and PCR assays developed for identification of H. annosum species Primers combination and sequence (50 –30 ) Specificity/Amplicon size Forward Reverse ITS1Fa (cttggtcatttagaggaagtaa) ITS4 Fungi H. annosum Am-P/518 bp ITS P1 (gtcggtcgggttcttttgatc) H. annosum Am-S/486 bp ITS S1 (gccgcgtcttctcacaaact) MLF (taaaaatttaaattagccataa) Mito7 (gccaatttattttgctacc) H. annosum s.s./230 bp Mito5 (taagaccgctatawaccagac) H. abietinum/195 bp H. parviporum/185 bp MLS (aaattagccatattttaaaag.) EfaHaFor (ctatgtcgcggtacagcttg) EfaHaRev (gcgaggayaagaagtaatcagca) H. annosum spp./169 bp H. annosum Am-P/71 bp EfaNAPFor (gtacatggtcactgtacgtagatgc) H. annosum s.s./69 bp EFAEuPFor (atggtcactgtacgtagatcatgc) MJ-F (ggtcctgtctggctttgc) MJ-R (ctgaagcacaccttgcca) H. annosum s.s./100 bp H. parviporum/350 bp KJ-F (ccattaacggaaccgacgtg) KJ-R (gtgcggctcattctacgctatc) H. annosum spp./400 bp HET-7 (cttctcacaaactcttcg) HET-8 (caggtcccccacaatcg) a Universal primer designed by Gardes and Bruns (1993) Target site References ITS Garbelotto et al. (1996) Mt-LrDNA Garbelotto et al. (1998); Gonthier et al. (2001, 2003) Gonthier et al. (2007) EFA IGS Unknown ITS Hantula and Vainio (2003) Bahnweg et al. (2002) G. Nicolotti et al. Advances in Detection and Identification of Wood Rotting Fungi 265 Table 12.5 Multiplex PCR assays developed by Guglielmo et al. (2008b) for identification of hazardous wood rotting fungi within Ganoderma, Inonotus and Phellinus Multiplex PCR Primers combination and sequence (50 -30 ) Specificity/ Amplicon size Target site Forward Reverse Mgano ITS1-F GadR (caggcaacaagtgcgctc) G. adspersum, G. pfeifferi, G. applanatum (from North ITS America)/211 bp GapR (gacacgcttcacaagctcc) G. applanatum (from Europe)/200 bp G. lucidum (from Europe)/193 bp GlR (ttcacgaagccccgcaag) G. resinaceum, G. lucidum (from North America)/178 bp GrR (aagagcccgcttcacaacg) Fomitiporia (P. punctatus, P. robustus)/258 bp Nuc-LrDNA Mhyme 25sF FomR (cccagcccatgtatacaatag) FuscR (cacactccgaagagtgcc) Fuscoporia (P. contiguus, P. gilvus, P.torulosus)/ 225 bp I. dryadeus/254 bp IdryaR (accgacgcatacaacaaagg) Inocutis (I. dryophilus)/265 bp InocuR (cctcagtccccgacggt) Inonotus s.s. (I. andersonii, I. hispidus, I. obliquus)/ InssR (gatgttgacccgtccgac) 214 bp PhssR (ggcgctacattccctctg) Phellinus s.s. (P. igniarius, P. tremulae, P. tuberculosus)/ 173 bp 12 Table 12.4 Multiplex PCR assays developed by Guglielmo et al. (2007) for identification of important wood rotting fungi hazardous for tree stability Specificity/Amplicon size Target Site Multiplex PCR Primers combination and sequence (50 -30 ) Forward Reverse M1 ITS1F ITS4 Fungi ITS Gano2R (tatagagtttgtgataaacgca) Ganoderma spp./226–228 bp Inonotus spp. and Phellinus spp./111 bp Nuc-LrDNA F115 (taagcgacccgtcttgaaac) Hyme2R (tgcdccccctygcggag) Hericium spp./199 bp Nuc-LrDNA M2 25sF (tggcgagagaccgatagc) Heri2R (cagcccttgtccggcagt) LaetR (ccgagcaaacgaatgcaa) L. sulphureus/146 bp Pleurotus spp./158 bp Pleu2R (aaccaggaagtacgcctcac) Armillaria spp./184 bp ITS ITS3 Armi2R (aaacccccataatccaatcc) M3 ITS3 PerR (atctgcaaagaccggtaaggt) P. fraxinea/152 bp ITS Schizophyllum spp./191 bp Schi2R (ctccagcagacctccacttc) Stereum spp./234–240 bp Ste2R (gtcgcaacaagacgcactaa) K. deusta/260 bp Ustu2Rb (gctcatctctacaggcgagaa) TraR (ttcatagtcttatggaaaccgc) Trametes spp./220 bp Mt-SrDNA MS1a (cagcagtcaagaatattagtcaatg) a Universal primer designed by White et al. 1990 b Primer designed by Nicolotti et al. (2009) 266 G. Nicolotti et al. Taxon-specific priming PCR is the major tool for detection and identification of wood rotting fungi. Indeed, this method is reproducible, specific, fast, easy to perform, and useful for analysis of environmental samples. However, a preliminary knowledge of the nucleotide sequence of target site is necessary, as well as a long work on primer design and testing for specificity and possible cross-reactions with other nontarget fungi. 12.5.6 Real-Time PCR Real-time PCR combines the conventional PCR with the generation of a fluorescent signal that depends on the amount of amplified DNA in each cycle. A detection system allows the measurement of this signal throughout the reaction, providing a real-time analysis and quantification of the specific DNA targets (Schmittgen 2001). The initial DNA amount of target DNA in the reaction can be related to a cycle threshold (Ct), defined as the cycle number at which there is a statistically significant increase of fluorescence. This method can thus be quantitative and does not require a further electrophoretic run to detect the amplicon. The most popular real-time PCR methods, such as Taq-man (Lee et al. 1993), Molecular beacons (Tyagi and Kramer 1996) and Scorpion-PCR (Whitcombe et al. 1999), are based on the use of a fluorescent reporter dye and a quencher linked to probes. These probes are designed to be specific to a complementary site in the amplicon. Real-time PCR has thus an increased specificity with respect to conventional PCR. A multiplex real-time PCR assay was developed to monitor the dynamics of Picea abies-H. annosum pathosystems (Hietala et al. 2003). In this study, real-time PCR proved to be more effective than traditional methods in screening clone resistance to the pathogen both under laboratory and field conditions. For diagnostic purposes, real-time PCRs were developed for R. necatrix and Fuscoporia torulosa (Pers.) T. Wagner & M. Fisch. (Campanile et al. 2008; Schena and Ippolito 2003). Real-time PCR allowed detection of R. necatrix both in infected plants and in contaminated soils with higher sensitivity than traditional isolation methods and baiting systems, respectively. Although real-time PCR methods are more rapid than conventional PCR and, thus being more appropriate for routinely diagnostics, the need of costly machines and reagents has limited so far their use in the field of diagnostics of wood rotting fungi. 12.5.7 DNA Sequencing As stated above, preliminary knowledge of nucleotide sequence of the target DNA region is paramount for the development of efficient diagnostic tools, such as 12 Advances in Detection and Identification of Wood Rotting Fungi 267 taxon-specific primers and RFLP markers. Additionally, the DNA sequence itself can also be used as a straightforward and powerful means of identification potentially exploiting polymorphism not only at a restricted site but everywhere in the DNA region considered. For this purpose, PCR product obtained from DNA extracts of the unknown sample is subjected, tout court or after a cloning step, to a sequencing reaction that leads, through high-resolution capillary analyzers, to a chromatogram displaying the complete nucleotide sequence for the DNA fragment. The unknown sequence is then compared, through Basic Local Alignment Search Tool (BLAST), to other sequences deposited in GenBank, which is an annotated database of all available nucleotide and amino acid sequences (Altschul et al. 1990). With the help of Expect (E) values, which report the significance of matches, it is possible to detect species displaying the highest sequence similarity with the unknown isolate. Large amounts of rDNA sequences, provided by several taxonomic and phylogenetic studies (Chillali et al. 1998; Hong and Jung 2004; Ko and Jung 1999; Larsson and Larsson 2003; Wagner and Fischer 2002), are increasingly available in GenBank for most of wood decay fungal taxa. The same is true for sequences of other nuc- and mt-DNA regions, such as ATP synthase subunit 6 (ATP), calmodulin (CAM), elongation factor 1-a (EFA), and glyceraldehyde 3-phosphate dehydrogenase (GPD) (Johannesson and Stenlid 2003; Linzer et al. 2008). Direct sequencing of ITS combined with BLAST search has proved to be effective and reliable for the identification of fungi directly from wood in construction (Högberg and Land 2004). The comprehensive database of ITS and ribosomal genes sequences provided by Moreth and Schmidt (2005) for the most important house rot species may be helpful to increase the chance of identification by means of BLAST analysis. Sequencing, after nested PCR, was helpful for the identification of H. annosum species in forests (Gonthier et al. 2003). Interestingly, sequencing was also helpful for the detection of an exotic root rot pathogen of forest trees in Italy (Gonthier et al. 2004). Sequence analysis of two nuclear loci and one mitochondrial locus showed that individuals from an Italian stone pine stand belonged to H. annosum Am-P (Gonthier et al. 2004). In a further study aimed at describing the patterns of invasion of this exotic pathogen, Gonthier et al. (2007) proved, through sequencing of the same loci, the occurrence of hybrids between North American and European Heterobasidion species. Finally, ITS sequencing of DNA extracted from both fruit bodies or decayed wood is a method currently used to investigate the fungi associated to root and heart rot of Acacia mangium (Glen et al. 2006). Although sequencing is very powerful for the identification of wood decay fungi, even directly from environmental samples after cloning, it appears too expensive as a routine analysis method. Furthermore, since in some cases species designation of the submitted organism in GenBank was wrong (Camacho et al. 1997; Redecker et al. 1999), a careful interpretation of the results from BLAST search is necessary. 268 Table 12.6 Comparison of diagnostic techniques for wood rotting fungi in terms of preliminary requirements and main features Diagnostic technique Requirement of: Specificity Reproducibility Cost Fruit bodies Isolation of Mycological background pure cultures Analysis of fruit bodies Yes No Yes High Medium Very low Analysis of pure fungal cultures No Yes Yes Medium Medium Low Protein-based techniques No Yes No High Low Medium Immunological techniques No No No Medium Medium Medium RAPD-PCR No Yes No High Low Low No High High Medium PCR-RFLP No Yesa T-RFLP No No No High High High Taxon-specific priming PCR No No No High High Low Real-Time PCR No No No High High High DNA-sequencing No Yesa No High High High Microarrays No No No High High High a Not required in case of preliminary cloning of PCR products Time required (days) 1–2 20–60 7–15 2–15 7–15 8–16 3–5 1–3 1–2 4–6 1–2 G. Nicolotti et al. 12 Advances in Detection and Identification of Wood Rotting Fungi 12.6 269 Conclusion and Perspectives The introduction of DNA-based identification tools has overcome several limits of traditional methods and protein-based methods, allowing for rapid fungal diagnosis even in the absence of unequivocal rot signs and directly from wood without the need of a time-consuming isolation step (Table 12.6). Through DNA-based techniques, identification of wood decay fungi can be performed at early stages of fungal infection, making possible effective timely treatments against harmful house rots and reducing risks of tree failures in urban environment. Further, most DNAbased techniques provide straightforward diagnostic characters easy to interpret, independently of factors related to environment and stage of fungal development. Advances in DNA-based phytodiagnostics are mainly addressed to the development of rapid methods allowing samples to be simultaneously screened for a large number of pathogens (Mumford et al. 2006). Multiplex taxon-specific PCRs have already allowed successful identification of several wood decay fungal species in a few assays (Guglielmo et al. 2007, 2008b). A more powerful tool for parallel testing of many targets in a single reaction is provided by microarrays, which are based on specific hybridization events between nucleic acid in a sample and known nucleic acid probes linked to a solid phase. A basic array method, the hybridization of immobilized sequence-specific oligonucleotide probes with PCR amplified fungal rDNA (SSOP) , has already been proved to be a valuable and sensitive tool for simultaneous detection of decay fungi involved in the deterioration of wood products in service (Oh et al. 2003). References Allmér J, Vasiliauskas R, Ihrmark K, Stenlid J, Dahlberg A (2006) Wood-inhabiting fungal communities in woody debris of Norway spruce (Picea abies (L.) Karst.), as reflected by sporocarps, mycelial isolations and T-RFLP identification. FEMS Microbiol Ecol 55:57–67 Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ (1990) Basic local alignment search tool. J Mol Biol 215:403–410 Amenduni T, Bazzoni A, Romanazzi G, Cariddi C, Vovlas N, Trisciuzzi N, Schena L, Potere O, Finetti-Sialer M, Myrta A (2001) Distribuzione dei patogeni delle drupacee in Puglia. Atti progetto PM A32 – Norme fitosanitarie e commercializzazione delle produzioni vivaistiche, Locorotondo, Bari, Italy, pp 143–179 Anderson JB (1986) Biological species of Armillaria in North America: redesignation of groups IV and VIII and enumeration of voucher strains for other groups. Mycologia 78:837–839 Anselmi N, Giorcelli A (1990) Factors influencing the incidence of Rosellinia necatrix Prill. in poplars. Eur J Forest Pathol 20:175–183 Anderson JB, Bailey SS, Pukkila PJ (1989) Variation in ribosomal DNA among biological species of Armillaria, a genus of root-infecting fungi. Evolution 43:1652–1662 Bahnweg G, Möller EM, Anegg S, Langebartels C, Wienhaus O, Sandermann H Jr (2002) Detection of Heterobasidion annosum s.l. [(Fr.) Bref.] in Norway spruce by polymerase chain reaction. J Phytopathol 150:382–389 270 G. Nicolotti et al. Bahnweg G, Schulze S, Moller EM, Rosenbrock H, Langebartels C, Sandermann H (1998) DNA isolation from recalcitrant materials such as tree roots, bark, and forest soil for the detection of fungal pathogens by polymerase chain reaction. Anal Biochem 262:79–82 Bech-Andersen J (1995) The dry rot fungus and other fungi in houses. Hussvamp Laboratoriet, Holte, Denmark Bernicchia A (2005) Polyporaceae s.l. Ed. Candusso, Alassio, Italy, p 808 Blanchette RA (1991) Delignification by wood-decay fungi. Annu Rev Phytopathol 29:381–398 Bravery AF, Berry RW, Carey JK, Cooper DE (2003) Recognising wood rot and insect damage in buildings, 2nd edn. Building Research Establishment, Watford Breitenbach J, Kränzlin F (1986) Champignons de Suisse, Champignons sans Lames, vol 2. Mykologia, Luzern, p 412 Bruns TD, Shefferson RP (2004) Evolutionary studies of ectomycorrhizal fungi: recent advances and future directions. Can J Bot 82:1122–1132 Camacho FJ, Gernandt DS, Liston A, Stone JK, Klein AS (1997) Endophytic fungal DNA, the source of contamination in spruce needle DNA. Mol Ecol 6:983–987 Campanile G, Schena L, Luisi N (2008) Real-time PCR identification and detection of Fuscoporia torulosa in Quercus ilex. Plant Pathol 57:76–83 Chase TE, Ullrich RC (1988) Heterobasidion annosum, root- and butt-rot of trees. Adv Plant Pathol 6:501–510 Chillali M, Wipf D, Guillaumin JJ, Mohammed C, Botton B (1998) Delineation of the European Armillaria species based on the sequences of the internal transcribed spacer (ITS) of ribosomal DNA. New Phytol 138:553–561 Clausen CA (1997) Immunological detection of wood decay fungi – an overview of techniques developed from 1986 to the present. Int Biodeter Biodegrad 39:133–143 Clausen CA, Kartal SN (2003) Accelerated detection of brown-rot decay: comparison of soil block test, chemical analysis, mechanical properties, and immunodetection. Forest Prod J 53:90–94 Deflorio G, Johnson C, Fink S, Schwarze FWMR (2008) Decay development in living sapwood of coniferous and deciduous trees inoculated with six wood decay fungi. Forest Ecol Manag 255:2373–2383 Delatour C (1980) Le Fomes annosus (Fr.) Cke. en Europe de l’ouest: importance economique, orientation des recherches. In: Dimitri L (Ed.) Proceedings of the 5th International Conference on problems of root and butt rots in conifers, August 7-12, 1978, Kassel, F.R.G., Germany, pp 9–18 Edel-Hermann W, Dreumont C, Perez-Piqueres A, Steinberg C (2004) Terminal restriction fragment length polymorphism analysis of ribosomal RNA genes to assess changes in fungal community structure in soils. FEMS Microbiol Ecol 47:397–404 Elnifro EM, Ashshi AM, Cooper RJ, Klapper PE (2000) Multiplex PCR: optimization and application in diagnostic virology. Clin Microbiol Rev 13:559–570 Erkkilä R, Niemelä T (1986) Polypores in the parks and forests of the City of Helsinki. Karstenia 26:1–40 Fischer M, Wagner T (1999) RFLP analysis as a tool for identification of lignicolous basidiomycetes: European polypores. Eur J Forest Pathol 29:295–304 Garbelotto M, Bruns TD, Cobb FW, Otrosina WJ (1993) Differentiation of intersterility groups and geographic provenances among isolates of Heterobasidion annosum detected by random amplified polymorphic DNA assays. Can J Bot 71:565–569 Garbelotto M, Otrosina WJ, Cobb FW, Bruns TD (1998) The European S and F intersterility groups of Heterobasidion annosum may represent sympatric protospecies. Can J Bot 76:397–409 Garbelotto M, Ratcliff A, Bruns TD, Cobb FW, Otrosina WJ (1996) Use of taxon-specific competitive-priming PCR to study host specificity, hybridization, and intergroup gene flow in intersterility groups of Heterobasidion annosum. Phytopathology 86:543–551 Gardes M, Bruns TD (1993) ITS primers with enhanced specificity for basidiomycetes – application to the identification of mycorrhizae and rusts. Mol Ecol 2:113–118 12 Advances in Detection and Identification of Wood Rotting Fungi 271 Germain H, Laflamme G, Bernier L, Boulet B, Hamelin RC (2002) DNA polymorphism and molecular diagnosis in Inonotus spp. Can J Plant Pathol 24:194–199 Gilbertson RL, Ryvarden L (1987) Megasporoporia –Wrightoporia, North American Polypores. Fungiflora, Oslo, Norway, pp 437–885 Glen M, Potter K, Sulistyawati P (2006) Molecular identification of organisms associated with root and heart rot in Acacia mangium. In: Potter, K., Rimbawanto, A. and Beadle, C. ed., Heart rot and root rot in tropical Acacia plantations. Canberra. ACIAR Proc 124:55–59 Göller K, Rudolph D (2003) The need for unequivocally defined reference fungi-genomic variation in two strains named as Coniophora puteana BAM Ebw. 15. Holzforschung 57:456–458 Gonthier P, Garbelotto M, Nicolotti G (2003) Swiss stone pine trees and spruce stumps represent an important habitat for Heterobasidion spp. in subalpine forests. Forest Pathol 33:191–203 Gonthier P, Garbelotto M, Varese GC, Nicolotti G (2001) Relative abundance and potential dispersal range of intersterility groups of Heterobasidion annosum in pure and mixed forests. Can J Bot 79:1057–1065 Gonthier P, Garbelotto MM, Nicolotti G (2005) Seasonal patterns of spore deposition of Heterobasidion species in four forests of the Western Alps. Phytopathology 95:759–767 Gonthier P, Nicolotti G (2007) A field key to identify common wood decay fungal species on standing trees. Arboric Urban For 33:410–420 Gonthier P, Nicolotti G, Linzer R, Guglielmo F, Garbelotto M (2007) Invasion of European pine stands by a North American forest pathogen and its hybridization with a native interfertile taxon. Mol Ecol 16:1389–1400 Gonthier P, Warner R, Nicolotti G, Mazzaglia A, Garbelotto MM (2004) Pathogen introduction as a collateral effect of military activity. Mycol Res 108:468–470 Guerin-Laguette A, Matsushita N, Kikuchi K, Iwase K, Lapeyrie F, Suzuki K (2002) Identification of a prevalent Tricholoma matsutake ribotype in Japan by rDNA IGS1 spacer characterization. Mycol Res 106:435–443 Guglielmo F (2005) A molecular approach for the detection and early identification of wood rotting fungi as useful tool in tree stability assessment, Ph.D. Thesis in Microbial Biotechnology, University of Florence, Florence, Italy, pp 131 Guglielmo F, Bergemann SE, Gonthier P, Nicolotti G, Garbelotto M (2007) A multiplex PCRbased method for the detection and early identification of wood rotting fungi in standing trees. J Appl Microbiol 103:1490–1507 Guglielmo F, Bergemann SE, Gonthier P, Nicolotti G, Garbelotto M (2008a) In: Garbelotto M, Gonthier, P (Eds.) Proceedings of the 12th International Conference on Root and Butt Rots of Forest Trees, Berkeley-California and Medford-Oregon, August 12–19, 2007. The University of California, Berkeley, USA, pp 196-200 Guglielmo F, Gonthier P, Garbelotto M, Nicolotti G (2008b) A PCR-based method for the identification of important wood rotting fungal taxa within Ganoderma, Inonotus s.l. and Phellinus s.l. FEMS Microbiol Lett 282:228–237 Guillaumin JJ, Lung B, Romagnesi H, Marxmüller H, Lamoure D, Durrieu G (1985) Systematics of the Armillaria mellea complex. Phytopathological consequences. Eur J Forest Pathol 15:268–277 Habermehl A, Ridder HW, Seidl P (1999) Computerized tomographic systems as tools for diagnosing urban tree health. Acta Hortic 496:261–268 Hansen EM, Goheen EM (2000) Phellinus weirii and other native root pathogens as determinants of forest structure and process in Western North America. Annu Rev Phytopathol 38:515–539 Hantula J, Vainio E (2003) Specific primers for the differentiation of Heterobasidion annosum (s. str.) and H. parviporum infected stumps in Northern Europe. Silva Fenn 37:181–187 Harrington TC, Wingfield BD (1995) A PCR-based identification method for species of Armillaria. Mycologia 87:280–288 Harrington TC, Worrall JJ, Rizzo DM (1989) Compatibility among host-specialized isolates of Heterobasidion annosum from western North America. Phytopathology 79:290–296 272 G. Nicolotti et al. Hickman GW, Perry EJ (1997) Ten common wood decay fungi on landscape trees-Identification handbook. Western Chapter, ISA, Sacramento Hietala AM, Eikenes M, Kvaalen H, Solheim H, Fossdal CG (2003) Multiplex real-time PCR for monitoring Heterobasidion annosum colonization in Norway spruce clones that differ in disease resistance. Appl Environ Microbiol 69:4413–4420 Hjortstam K, Larsson KH, Ryvarden L (1978) The corticiaceae of North Europe, vol 1. Fungiflora, Oslo, Norway, p 60 Högberg N, Land CJ (2004) Identification of Serpula lacrymans and other decay fungi in construction timber by sequencing of ribosomal DNA – a practical approach. Holzforschung 58:199–204 Hong S, Jeong W, Jung H (2002) Amplification of mitochondrial small subunit ribosomal DNA of polypores and its potential for phylogenetic analysis. Mycologia 94:823–833 Hong S, Jung H (2004) Phylogenetic analysis of Ganoderma based on nearly complete mitochondrial small-subunit ribosomal DNA sequences. Mycologia 96:742–755 Huckfeldt T, Schmidt O (2006) Identification key for European strand-forming house-rot fungi. Mycologist 20:42–56 Intini M, Panconesi A, Parrini C (2000) Malattie delle Alberature in Ambiente Urbano. CNRIPAF, Florence, Italy, p 216 Jahnke KD, Bahnweg G, Worrall JJ (1987) Species delimitation in the Armillaria mellea complex by analysis of nuclear and mitochondrial DNAs. Trans Br mycol Soc 88:572–575 Jasalavich CA, Ostrofsky A, Jellison J (2000) Detection and identification of decay fungi in spruce wood by restriction fragment length polymorphism analysis of amplified genes encoding rRNA. Appl Environ Microbiol 66:4725–4734 Jellison J, Goodell B (1988) Immunological detection of decay in wood. Wood Sci Technol 22:293–297 Jellison J, Goodell B (1989) Inhibitory effects of undecayed wood and the detection of Postia placenta using the enzyme-linked immunosorbent assay. Wood Sci Technol 23:13–20 Jellison J, Howell C, Schilling J, Goodell B, Quarles S (2004) Investigations into the biology of Meruliporia incrassata. In: International Research Group on Wood Preservation, IRG/WP Series Document 04-10508 Johannesson H, Stenlid J (1999) Molecular identification of wood-inhabiting fungi in an unmanaged Picea abies forest in Sweden. Forest Ecol Manage 115:203–211 Johannesson H, Stenlid J (2003) Molecular markers reveal genetic isolation and phylogeography of the S and F intersterility groups of the wood-decay fungus Heterobasidion annosum. Mol Phylogenet Evol 29:94–101 Karjalainen R (1996) Genetic relatedness among strains of Heterobasidion annosum as detected by random amplified polymorphic DNA markers. J Phytopathol 144:399–404 Kasuga T, Mitchelson KR (2000) Intersterility group differentiation in Heterobasidion annosum using ribosomal IGS1 region polymorphism. Forest Pathol 30:329–344 Kasuga T, Woods C, Woodward S, Mitchelson K (1993) Heterobasidion annosum 5.8s ribosomal DNA and internal transcribed spacer sequence: rapid identification of European intersterility groups by ribosomal DNA restriction polymorphism. Curr Genet 24:433–436 Kennedy N, Clipson N (2003) Fingerprinting the fungal community. Mycologist 17:158–164 Khairudin H (1995) Basal stem rot of oil palm caused by Ganodema boninense. In: PORIM International Palm Oil Congress: update and vision (Agriculture), 20–25 September 1993, Kuala Lumpur, Malaysia, pp 739–749 Kile GA, McDonald GI, Byler JW (1991) Ecology and disease in natural forests. In: Shaw CG, Kile GA (eds) Armillaria root disease. Washington D.C, USDA Forest Service, pp 102–121 Kirk PM, Cannon PF, David JC, Stalpers J (2001) Ainsworth and Bisby’s dictionary of the fungi, 9th edn. CAB International, Wallingford, UK, p 655 Kleist G, Seehann G (1999) Der Eichenporling, Donkioporia expansa, ein wenig bekannter Holzzerstörer in Gebäuden. Z Mykol 65:23–32 12 Advances in Detection and Identification of Wood Rotting Fungi 273 Ko K, Jung H (1999) Phylogenetic re-evaluation of Trametes consors based on mitochondrial small subunit ribosomal DNA sequences. FEMS Microbiol Lett 170:181–186 Korhonen K (1978) Intersterility groups of Heterobasidion annosum. Commun Inst Forest Fenn 94:1–25 Korhonen K, Stenlid J (1998) Biology of Heterobasidion annosum. In: Woodward S, Stenlid J, Karjalainen R, Hüttermann A (eds) Heterobasidion annosum: biology, ecology, impact and control. CAB International, Wallingford, Oxon, UK, pp 43–70 Larsson E, Larsson KH (2003) Phylogenetic relationships of russuloid basidiomycetes with emphasis on aphyllophoralean taxa. Mycologia 95:1037–1065 Lee LG, Connell CR, Bloch W (1993) Allelic discrimination by nick-translation PCR with fluorogenic probes. Nucleic Acids Res 21:3761–3766 Lee SS (2000) The current status of root diseases of Acacia mangium Willd. Ganoderma diseases of perennial crops. CAB International, Wallingford, Oxon, UK, pp 71–79 Lim YW, Yeung YCA, Sturrock R, Leal I, Breuil C (2005) Differentiating the two closely related species, Phellinus weirii and P. sulphurascens. Forest Pathol 35:305–314 Linzer RE, Otrosina WJ, Gonthier P, Bruhn J, Laflamme G, Bussieres G, Garbelotto M (2008) Inferences on the phylogeography of the fungal pathogen Heterobasidion annosum, including evidence of interspecific horizontal genetic transfer and of human-mediated, long-range dispersal. Mol Phylogenet Evol 46:844–862 Liu WT, Marsh TL, Cheng H, Forney LJ (1997) Characterization of microbial diversity by determining terminal restriction fragment length polymorphisms of genes encoding 16S rRNA. Appl Environ Microbiol 63:4516–4522 Lochman J, Sery O, Mikes V (2004) The rapid identification of European Armillaria species from soil samples by nested PCR. FEMS Microbiol Lett 237:105–110 Lombard FF, Chamuris GP (1990) Basidiomycetes. In: Wang CJK, Zabel RA (eds) Identification manual for fungi from utility poles in the eastern United States. American Type Culture Collection, Rockville, USA, pp 21–104 Lonsdale D (1999) Principles of tree hazard assessment and management. Forestry Commission, London, UK, pp 388 Matsushita N, Suzuki K (2005) Identification of Armillaria species in Japan using PCR-RFLP analysis of rDNA intergenic spacer region and comparisons of Armillaria species in the world. J Forest Res 10:173–179 Mattheck C, Breloer H (1992) Tree monitoring with VTA – visual tree assessment. Baumkontrollen mit VTA visual tree assessment. Gartenamt 41:777–784 Mattheck C, Breloer H (1995) The body language of trees: a handbook for failure analysis, Research for Amenity Trees 4. HMSO, London, p 240 Moreth U, Schmidt O (2000) Identification of indoor rot fungi by taxon-specific priming polymerase chain reaction. Holzforschung 54:1–8 Moreth U, Schmidt O (2005) Investigations on ribosomal DNA of indoor wood decay fungi for their characterization and identification. Holzforschung 59:90–93 Möykkynen T, Miina J (2002) Optimizing the management of a butt-rotted Picea abies stand infected by Heterobasidion annosum from the previous rotation. Scand J Forest Res 17:47–52 Müller U, Bammer R, Halmschlager E, Stollberger R, Wimmer R (2001) Detection of fungal wood decay using magnetic resonance imaging. Holz Roh Werkst 59:190–194 Mumford R, Boonham N, Tomlinson J, Barker I (2006) Advances in molecular phytodiagnostics – new solutions for old problems. Eur J Plant Pathol 116:1–19 Nicolotti G, Gonthier P (2005) Stump treatment against Heterobasidion with Phlebiopsis gigantea and some chemicals in Picea abies stands in the western Alps. Forest Pathol 35:365–374 Nicolotti G, Gonthier P, Guglielmo F, Garbelotto M (2009) A biomolecular method for the detection of wood decay fungi: a focus on tree stability assessment. Arboric Urban For 35:14–19 Nicolotti G, Gonthier P, Pecollo D (2004a) Ecologia e grado di preferenza d’ospite dei funghi agenti di carie/I parte. Acer 1:47–51 274 G. Nicolotti et al. Nicolotti G, Gonthier P, Pecollo D (2004b) Ecologia e grado di preferenza d’ospite dei funghi agenti di carie/II parte. Acer 2:59–67 Nicolotti G, Socco LV, Martinis R, Godio A, Sambuelli L (2003) Application and comparison of three tomographic techniques for detection of decay in trees. J Arboric 29:66–78 Nobles MK (1965) Identification of cultures of wood-inhabiting hymenomycetes. Can J Bot 43:1097–1139 Oh SK, Kamdem DP, Keathley DE, Han KH (2003) Detection and species identification of wooddecaying fungi by hybridization of immobilized Sequence-Specific Oligonucleotide Probes with PCR-amplified fungal ribosomal DNA internal transcribed spacers. Holzforschung 57:346–352 Otrosina WJ, Chase TE, Cobb FW Jr (1992) Allozyme differentiation of intersterility groups of Heterobasidion annosum isolated from conifers in the western United States. Phytopathology 82:540–545 Otrosina WJ, Chase TE, Cobb FW Jr, Korhonen K (1993) Population structure of Heterobasidion annosum from North America and Europe. Can J Bot 71:1064–1071 Palfreyman JW, Bruce A, Button D, Glancy H, Vigrow A, King B (2001) Immunological methods for the detection and characterisation of wood decay basidiomycetes. Int Biodeter Biodegr 48:74–78 Palfreyman JW, Vigrow A, King B (1991) Molecular identification of fungi causing rot of building timbers. Mycologist 5:73–77 Pegler DN (2000) Taxonomy, nomenclature and description of Armillaria. In: Fox RTV (ed) Armillaria root rot: biology and control of honey fungus. Intercept, Andover, pp 81–93 Piri T (1996) The spreading of the S type of Heterobasidion annosum from Norway spruce stumps to the subsequent tree stand. Eur J Forest Pathol 26:193–204 Priestley R, Mohammed C, Dewey FM (1994) The development of monoclonal antibody-based ELISA and dipstick assays for the detection and identification of Armillaria species in infected wood. In: Schots A, Dewey FM, Oliver RP (eds) Modem assays for plant pathogenic fungi. CAB International, Oxford, pp 149–156 Råberg U, Högberg NOS, Land CJ (2005) Detection and species discrimination using rDNA T-RFLP for identification of wood decay fungi. Holzforschung 59:696–702 Rayner ADM, Boddy L (1988) Fungal decomposition of wood. Its biology and ecology. Wiley, Chichester, pp 587 Redecker D, Hijri M, Dulieu H, Sanders IR (1999) Phylogenetic analysis of a dataset of fungal 5.8S rDNA sequences shows that highly divergent copies of internal transcribed spacers reported from Scutellospora castanea are of Ascomycete origin. Fungal Genet Biol 28:238–244 Rizzo DM, Harrington TC (1993) Delineation and biology of clones of Armillaria ostoyae, A. gemina and A. calvescens. Mycologia 85:164–174 Schena L, Ippolito A (2003) Rapid and sensitive detection of Rosellinia necatrix in roots and soils by real time Scorpion-PCR. J Plant Pathol 85:15–25 Schmidt O (2007) Indoor wood-decay basidiomycetes: damage, causal fungi, physiology, identification and characterization, prevention and control. Mycol Prog 6:261–279 Schmidt O, Huckfeldt T (2005) Gebäudepilze. In: Müller J (ed) Holzschutz im Hochbau. Fraunhofer IRB, Stuttgart, pp 44–72 Schmidt O, Moreth U (1995) Detection and differentiation of Poria indoor brown-rot fungi by polyacrylamide gel electrophoresis. Holzforschung 49:11–14 Schmidt O, Moreth U (1998) Characterization of indoor rot fungi by RAPD analysis. Holzforschung 52:229–233 Schmidt O, Moreth U (1999) Identification of the dry rot fungus, Serpula lacrymans, and the wild merulius, S. himantioides, by amplified ribosomal DNA restriction analysis (ARDRA). Holzforschung 53:123–128 Schmidt O, Moreth U (2000) Species-specific PCR primers in the rDNA-ITS region as a diagnostic tool for Serpula lacrymans. Mycol Res 104:69–72 12 Advances in Detection and Identification of Wood Rotting Fungi 275 Schmittgen TD (2001) Real-time quantitative PCR. Methods 25:383–385 Schulze S, Bahnweg G, Möller EM, Sandermann H Jr (1997) Identification of the genus Armillaria by specific amplification of an rDNA-ITS fragment and evaluation of genetic variation within A. ostoyae by rDNA-RFLP and RAPD analysis. Eur J Forest Pathol 27:225–239 Schwarze FWMR, Baum S (2000) Mechanisms of reaction zone penetration by decay fungi in wood of beech (Fagus sylvatica). New Phytol 146:129–140 Schwarze FWMR, Engels J, Mattheck C (2004) Fungal strategies of wood decay in trees, 2nd edn. Springer, Berlin Sicoli G, Fatehi J, Stenlid J (2003) Development of species-specific PCR primers on rDNA for the identification of European Armillaria species. Forest Pathol 33:287–297 Sierra AP, Whitehead DS, Whitehead MP (1999) Investigation of a PCR-based method for the routine identification of British Armillaria species. Mycol Res 103:1631–1636 Smith ML, Bruhn JN, Anderson JB (1992) The fungus Armillaria bulbosa is among the largest and oldest living organisms. Nature 356:428–431 Stalpers JA (1978) Identification of wood-inhabiting fungi in pure culture, Stud Mycol 16. Centraalbureau Schimmelcultures, Baarn, p 248 Strouts RG, Winter TG (1994) Diagnosis of ill-health in trees. Forestry Commission, London, UK, pp 308 Schulze S, Bahnweg G, Tesche M, Sandermann H Jr (1995) Identification of European Armillaria species by restriction-fragment length polymorphisms of ribosomal DNA. Eur J For Pathol 25:214–223 Suhara H, Maekawa N, Kubayashi T, Kondo R (2005) Specific detection of a basidiomycete, Phlebia brevispora associated with butt rot of Chamaecyparis obtusa, by PCR-based analysis. J Wood Sci 51:83–88 Suhara H, Maekawa N, Kubayashi T, Sakai K, Kondo R (2002) Identification of the basidiomycetous fungus isolated from butt rot of the Japanese cypress. Mycoscience 43:477–481 Swiecki TJ, Bernhardt E, Drake C, Costello LR (2005) Relationships between Phytophthora ramorum canker (sudden oak death) and failure potential in coast live oak. In: Proceedings of the Sudden Oak Death Second Science Symposium, 19–21 January, 2005; Monterey. Pacific Southwest Research Station, Forest Service, U.S. Department of Agriculture, Albany, CA. Gen. Tech. Rep. PSW-GTR-196, pp 427–453 Terho M, Hantula J, Hallaksela AM (2007) Occurrence and decay patterns of common wooddecay fungi in hazardous trees felled in the Helsinki City. Forest Pathol 37:420–432 Tomikawa Y, Iwase Y, Arita K, Yamada H (1990) Nondestructive inspection of wooden poles using ultrasonic computed tomography. IEEE Trans UFFC 33:354–358 Tsai JN, Hsieh WH, Ann PJ, Yang CM (2007) Development of specific primers for Phellinus noxius. Plant Pathol Bull 16:193–202 Tyagi S, Kramer FR (1996) Molecular beacons: probes that fluoresce upon hybridization. Nat Biotechnol 14:303–308 Ullrich RC, Anderson JB (1978) Sex and diploidy in Armillaria mellea. Exp Mycol 2:119–129 Utomo C, Niepold F (2000) Development of diagnostic methods for detecting Ganodermainfected oil palms. J Phytopathol 148:507–514 Utomo C, Werner S, Niepold F, Deising HB (2005) Identification of Ganoderma, the causal agent of basal stem rot disease in oil palm using a molecular method. Mycopathologia 159:159–170 Vainio EJ, Hantula J (2000) Direct analysis of wood-inhabiting fungi using denaturing gradient gel electrophoresis of amplified ribosomal DNA. Mycol Res 104:927–936 Wagner T, Fischer M (2002) Proceedings towards a natural classification of the worldwide taxa Phellinus s.l. and Inonotus s.l., and phylogenetic relationships of allied genera. Mycologia 94:998–1016 Wahlstrom K, Karlsson JO, Holdenrieder O, Stenlid J (1991) Pectinolytic activity and isozymes in European Armillaria species. Can J Bot 69:2732–2739 Wargo PM, Harrington TC (1991) Host stress and susceptibility. In: Shaw CG, Kile GA (eds) Armillaria root disease. Washington D.C, USDA Forest Service, pp 88–101 276 G. Nicolotti et al. Watling R, Kile GA, Burdsall HHJ (1991) Nomenclature, taxonomy, and identification. In: Shaw CG, Kile GA (eds) Armillaria root disease. USDA Forest Service, Washington D.C., pp 1–9 Whitcombe D, Theaker J, Guy SP, Brown T, Little S (1999) Detection of PCR products using selfprobing amplicons and fluorescence. Nat Biotechnol 17:804–807 White TJ, Bruns TD, Lee S, Taylor JW (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ (eds) PCR protocols: a guide to methods and applications. Academic press, San Diego, USA, pp 315–322 Williams JGK, Kubelik AR, Livak KJ, Rafalski JA, Tingey SV (1990) DNA polymorphisms amplified by arbitrary primers are useful as genetic markers. Nucleic Acids Res 18:6531–6535 Woodward S, Stenlid J, Karjalainen R, Hüttermann A (Eds.) (1998) Preface. In: Heterobasidion annosum, biology, ecology, impact and control. CAB International, Oxon, UK, pp xi–xii Worrall JJ, Anagnost SE, Zabel RA (1997) Comparison of wood decay among diverse lignicolous fungi. Mycologia 89:199–219 Chapter 13 Molecular Diversity and Identification of Endophytic Fungi Liang-Dong Guo Abstract Endophytes are organisms inhabiting the living plant organs at some time in their life, without causing apparent harm to the host. Endophytic fungi, which have been widely studied in various geographical and climatic zones, are ubiquitous and occur within all examined plants including a broad range of host orders, families, genera, and species in diverse ecosystems. DNA fingerprinting and sequencing techniques employed in the population genetic diversity and in the detection and identification of endophytic fungi are summarized in this chapter. 13.1 Introduction The term “endophyte,” originally introduced by De Bary (1866), refers to any organisms occurring within plant tissues, distinct from the epiphytes that live on plant surfaces. Carroll (1986) defines endophytes as mutualists, those fungi that colonize aerial parts of living plant tissues and do not cause symptoms of disease. Pathogenic and mycorrhizal fungi are excluded from this definition. Petrini (1991) considers that endophytes are organisms inhabiting the living plant organs at some time in their life, without causing apparent harm to the host. Therefore, latent pathogens known to live symptomlessly inside the host tissues and organisms that have an epiphytic phase in their life cycle are also endophytes. This latter definition is broad enough to include virtually any microbes and vascular plants that colonize the living internal tissues of plants (Bills 1996; Stone et al. 2000; Schulz and Boyle 2005, 2006). However, mycologists have come to employ this term “endophyte” (or endophytic fungi and fungal endophyte) only for those fungi that colonize a plant L.‐D. Guo Systematic Mycology & Lichenology Laboratory, Institute of Microbiology, Chinese Academy of Sciences, Beijing 100101, China e-mail: guold@sun.im.ac.cn Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_13, # Springer-Verlag Berlin Heidelberg 2010 277 278 L.-D. Guo without causing visible disease symptoms at any specific moment (Petrini 1991; Wilson 1995; Stone et al. 2000; Schulz and Boyle 2005). Endophytic fungi, which have been widely studied in various geographical and climatic zones, are ubiquitous and occur within all examined plants including a broad range of host orders, families, genera, and species in diverse ecosystems (Petrini 1991; Wilson 1995; Stone et al. 2000; Guo 2001; Schulz and Boyle 2005; Li et al. 2007; Sun and Guo 2007; Wei et al. 2007; Guo et al. 2008; Sun et al. 2008). Most endophytic fungi are members of the Ascomycota or their mitosporic fungi but can also include some taxa of the Basidiomycota, Zygomycota, and Oomycota (Zheng and Jiang 1995; Sinclair and Cerkauskas 1996). Because the plant tissues are multilayered and spatially and temporally diverse microbial habitats, they support a rich and varied endophytic mycobiota that form specialized association with various plant species and tissues. Consequently, an accepted estimate of 1.5 million fungal species exists on earth primarily based on the ratios of vascular plants to fungal species at 1:6, but endophytic fungi have not been seriously taken into account in the estimation (Hawksworth 1991). Furthermore, Petrini (1991) suggests that there should be more than 1 million species of endophytic fungi remaining to be discovered and described in the world based on the ratios of vascular plants to fungal species at 1:4–5. In the survey of endophytic fungal diversity, many techniques have been used; traditional cultivation-dependent techniques, however, have been routinely employed in previous studies. In the cultivation-dependent methods, living plant tissues are subjected to a serial process of surface sterilization in order to remove all organisms from the surface of plant tissues. Only internal fungi are isolated by means of the incubation of the plant samples onto nutrient plates. The cultivationdependent techniques have generally involved three basic steps: (1) surface sterilization of plant tissues to kill any fungi on the host surface, (2) isolation of endophytic fungi growing out from samples placed onto nutrient agar, and (3) identification of the endophytic fungi based on morphological characteristics in culture. The advantage of the cultivation-dependent method is that this technique is effective for rapid recovery of a large number of endophytic fungal species from plant tissues. However, the study of endophytes is a method-dependent process. Endophytic fungal communities obtained from plants are directly affected by surface sterilization techniques and incubation conditions, and whether the isolates sporulate. Therefore, there are limitations in the cultural isolation techniques: (1) It is rather laborious and time intensive and is not suitable to compare large numbers of samples; (2) The large number of sterile isolates poses a special problem, because they cannot be identified to any taxonomic category, while various methods have been used to promote sporulation of isolates in order to overcome the shortcomings of some isolates failing to sporulate in culture (Taylor et al. 1999; Guo et al. 1998, 2000, 2008); (3) Some fungi may be missed as a result of failure to grow or some grow slowly and are easily outcompeted by fast-growing species in artificial conditions. In order to overcome the potential technical bias, cultivationindependent approaches, e.g., molecular techniques, to analyze endophytic fungal communities of plants are needed. 13 Molecular Diversity and Identification of Endophytic Fungi 279 Molecular techniques have been successfully used in the detection and identification of mycorrhizal fungi in roots and soils (Gardes et al. 1991; Simon et al. 1993; Clapp et al. 1995; Chelius and Triplett 1999; Tedersoo et al. 2008); pathogenic fungi directly from within plant tissues (Schesser et al. 1991; Mills et al. 1992; Moukhamedov et al. 1994; Beck and Ligon 1995; Bates et al. 2001; Atkins et al. 2003, 2004); and fungi in Iceman’s grass clothing (Rollo et al. 1995), bamboos (Zhang et al. 1997), and glacial ice strata (Ma et al. 1997). In this chapter, molecular techniques, i.e., DNA fingerprinting and sequencing methods, employed in the population genetic diversity and in the detection and identification of endophytic fungi, excluding mycorrhizal fungi, are briefly summarized. 13.2 Molecular Fingerprinting for Endophytic Fungal Population DNA fingerprinting techniques, such as restriction fragment length polymorphism (RFLP), terminal-RFLP (T-RFLP), random amplified polymorphic DNA (RAPD), simple sequence repeat (SSR) or inter-SSR (ISSR), amplified fragment length polymorphism (AFLP), denaturing gradient gel electrophoresis (DGGE), temperature gradient gel electrophoresis (TGGE), and single-stranded conformation polymorphism (SSCP), are well established and have been successfully applied to assess the population genetic diversity and fungal communities in natural environment (Gardes et al. 1991; Simon et al. 1993; Chambers et al. 1999; Smit et al. 1999; Tooley et al. 2000; van Elsas et al. 2000; Bock et al. 2002; Klamer et al. 2002; Anderson et al. 2003; Huai et al. 2003; Jansa et al. 2003; Anderson and Cairney 2004; Liang et al. 2004, 2005). These techniques have recently been adopted and applied to assess the population structure and community of endophytic fungi. 13.2.1 RAPD and RFLP Techniques The RAPD technique is used to investigate the genotypic diversity in populations of an endophytic fungus Rhabdocline parkeri isolated from Douglas fir growing in various habitats (McCutcheon and Carroll 1993). A significantly lower number of R. parkeri genotypes per unit foliage have been isolated from trees within a 20-year-old managed stand and from an isolated tree than from old growth trees. Therefore, the variation of genetic diversity of R. parkeri populations is ascribed to differences in tree age and access to inoculum. Furthermore, the genotypes of an endophytic fungus Discula umbrinella are related to host origin on the basis of the analysis of 30 strains isolated from beech, chestnut, and oak assessed by RAPD markers (Hämmerli et al. 1992). The population structure of an endophytic fungus Phialocephala fortinii has been studied at a primary succession site on a glacier forefront using RAPD markers, and 23 genets of P. fortinii were detected in 34 strains in 1 year, and 10 genets were found in 49 strains in the next year, but none of 280 L.-D. Guo the genets was isolated in both years (Jumpponen 1999). Further studies show that there are recombination, gene, and genotype flow in P. fortinii population based on the analysis of single-locus RFLP markers (Grünig et al. 2003). In addition, all strains of endophytic Phyllosticta species isolated from different tropical tree species in India have been identified as species Phyllosticta capitalensis using ITS-RFLP markers (Pandey et al. 2003). 13.2.2 SSR and AFLP Techniques SSR or ISSR technique, which is originally used to measure genetic diversity of plants and animals (Zietkiewicz et al. 1994), has been applied in studies of fungi (Hantula et al. 1996; Hantula and Müller 1997; Liang et al. 2005). In endophyte studies, Groppe et al. (1995) have analyzed the genetic diversity of an endophytic fungus Epichloe¨ typhina isolated from tissues of Bromus erectus using a microsatellite-containing locus as molecular markers. Further, Groppe and Boller (1997) have developed specific primer pairs flanking a microsatellite-containing locus and successfully detected a rDNA fragment of endophytic Epichloe¨ species from infected tissues of B. erectus, but no fragments were generated from total DNA isolated from uninfected plant material or unrelated fungi isolated from the same grass. Further studies have shown that there are high levels of polymorphism between Neotyphodium and Epichloe¨ species and low levels of polymorphism within Neotyphodium coenophialum and N. lolii based on the analysis of polymorphic SSR markers, and these markers can be used to identify endophytic fungi Neotyphodium and Epichloe¨ and to evaluate intraspecific population genetic diversity (De Jong et al. 2003). Four different morphotypes of an endophytic fungus Sphaeropsis sapinea have been isolated from the natural and exotic Pinus spp. in the Southern Hemisphere (Burgess et al. 2001). Of these morphotypes, the putative I is found to be identical to Botryosphaeria obtusa, the other remaining three are clearly distinguished using polymorphic SSR markers. Endophytic fungi P. fortinii and type I of a nonsporulating mycelium, which have the same allozyme phenotype, were differentiated on the basis of ISSR analysis (Grünig et al. 2001). Furthermore, 21 genets were detected in 144 P. fortinii strains isolated from roots of Norway spruce (Picea abies) collected within a plot (3  3 m2) of a 40-year-old plantation using ISSR markers (Grünig et al. 2002). Further population genetic studies suggest that the endophytic fungus P. fortinii population has high genetic diversity and should be considered cryptic species in the same forest site and even in the same root fragment in Europe (Grünig et al. 2004, 2006, 2007, 2008). The relationship between endophytic population genotypes and hosts, age, and geographic origin has been investigated on the basis of the SSR analysis. A high genetic diversity in an endophytic fungus Alternaria alternata population isolated from Pinus tabulaeformis in Beijing was detected and no relationship between genotypes of A. alternata and host tissue ages (Guo et al. 2004) was found. Similarly, the genotypes of an endophytic fungus Guignardia mangiferae do not 13 Molecular Diversity and Identification of Endophytic Fungi 281 correspond either to the host or to the geographic origin (Rodrigues et al. 2004), and there is no host specificity for isolates of an endophytic fungus Lasiodiplodia theobromae, although there is very high gene flow between populations from different hosts based on the SSR analysis (Mohali et al. 2005). AFLP markers were used to analyze the genetic polymorphism existing in two natural populations of an endophytic fungus Epichloe¨ festucae in semiarid natural grasslands in western Spain, and most genetic variation detected was found to occur within populations, with only a moderate amount of genetic differentiation between populations, and nonrecombinant asexual reproduction predominated in both populations (Garcı́a et al. 2002). The SSR (or ISSR) technique is comparatively cheap, fast, and easy to perform. It is similar to RAPD analysis, but longer primers (ca. 18 nucleotides) are used and the conditions (e.g., annealing temperature) during amplification are more stringent. Furthermore, genomic regions containing microsatellites evolve and mutate more rapidly than other areas of genome. This is due to slipped-strand mispairing during replication, with the slippage rate depending on the length of the repeat (Levinson and Gutman 1987; Burgess et al. 2001). Therefore, the use of higher annealing temperatures and longer nucleotide primers results in highly reproducible SSR markers that are much more robust than the RAPD markers used previously (Roberts et al. 2000; Peever et al. 2002; Liang et al. 2005). The SSR markers are also more powerful than the RFLP profiles generated from rDNA in revealing genetic variation among a set of closely related isolates (Adachi et al. 1993; Aradhya et al. 2001). Thus, SSR (ISSR) technique combines most of the benefits of RAPD and microsatellite analyzes, and is ideal for studies of genetic variation of endophyte population. There are arguments against using SSR (ISSR) techniques, as compared to AFLP and RAPD, in population genetic studies. Although microsatellite alleles are considered to be codominant markers, differences in alleles are measured solely on the basis of size. There is therefore the possibility of single-point mutations within the flanking sequence that do not result in a change in the fragment length. Furthermore, fragments from different genomic regions can co-migrate because they are of the same size. It is possible that different indels could result in fragments of the same size that have different sequences. In addition, markers may not be independent, because of genetic linkage or being alternative alleles at the same locus. In an asexual fungus, however, meiotic segregation of markers cannot occur, and although co-migration may occur, this does not negate the usefulness of this approach. Thus, a dominant marker system is suitable for assessing haploid, asexual populations without overestimating variation due to co-segregation (Bock et al. 2002). 13.2.3 DGGE Technique The DGGE technique, which is capable of separating closely related sequences by their differential mobilities in a gradient of denaturants, has been effectively used to estimate the diversity of prokaryotes and eukaryotes in natural samples 282 L.-D. Guo (Dı́ez et al. 2001; Dar et al. 2005; Countway et al. 2005; Jeewon and Hyde 2006). This technique has recently been successfully applied to document fungal communities (Kowalchuk et al. 1997; Vainio and Hantula 2000; May et al. 2001; Nikolcheva et al. 2003). In endophyte studies, Duong et al. (2006) used DGGE coupled with sequence analysis of partial 18S rRNA gene to assess endophytic fungal diversity in living leaves of Magnolia liliifera collected from Thailand. A total of 14 operational taxonomic units were recovered, and the DGGE could be used to detect known and abundant fungi (Xylariales, Hypocreales, and Pleosporales) as well as unknown endophytic fungi (Mycosphaerellales, Dothideales, Helotiales, and Rhytismatales). Similarly, the composition and relative abundance of endophytic fungi were assessed by DGGE analysis of 18S rRNA gene fragments amplified from total community DNA extracted from roots of potato Solanum tuberosum (Götz et al. 2006). Dominant bands in DGGE correspond to Verticillium dahliae, Cylindrocarpon destructans, and Colletotrichum coccodes, as the most frequently isolated species by traditional cultural method. Therefore, differences in the relative abundance of endophytic fungi colonizing the roots of T4-lysozyme producing potatoes and the parental line can be detected by DGGE methods. DGGE is a suitable method that can be applied to estimate fungal diversity by excising and sequencing bands, thereby obtaining taxonomic information for members of the community via database searches and phylogenetic analysis (Anderson and Cairney 2004; Duong et al. 2006). Simultaneously, the techniques can be used in conjunction with DNA oligonucleotide probes to increase the specificity of the analysis (Stephen et al. 1998). Despite the advantages of DGGE, there are also disadvantages. In general, shorter fragments (<500 bp) of DNA result in better resolution between bands in a profile, thereby limiting the taxonomic information to properly identify taxa at the genus or species level (May et al. 2001; Duong et al. 2006), although some larger products have also been used successfully in a few cases (Ranjard et al. 2000). Moreover, even the most sensitive staining methods are often not sensitive enough to detect all the diversity present within a sample, particularly for the less dominant members of the fungal community (Anderson and Cairney 2004). In addition, in some cases single bands on a gel have been shown to comprise more than a single sequence type (Schmalenberger and Tebbe 2003). Further studies should consider primers that are more universal for fungi and give better phylogenetic resolution at generic or species level. 13.3 Molecular Sequencing for Endophytic Fungal Community Molecular fingerprinting techniques are inadequate for the analysis of fungal communities from environmental samples where several different fungi may be simultaneously present and where their identities unknown. However, molecular sequencing techniques have been successfully employed for fungal identification and phylogenies based on the sequence analyses of coding genes, e.g., cytochrome c oxidase 1 (CO1) gene, beta-tubulin 2 gene (tub2), and 18S, 28S and 5.8S genes of 13 Molecular Diversity and Identification of Endophytic Fungi 283 rDNA and noncoding internal transcribed spacer (ITS) regions of rDNA. Because most coding genes are highly conserved, they have been successfully used to assess phylogenetic relationships at higher taxonomic levels. The CO1, tub2, and ITS regions benefit from a fast rate of evolution, resulting in greater sequence variation between closely related species. These region sequences therefore generally provide greater lower taxonomic resolution at genus and species level. Molecular sequencing techniques have recently been successfully used in the detection and identification of endophytic fungi based on phylogenetic analysis and sequence similarity comparison. 13.3.1 Identification of Endophytic Fungi The endophytic fungal community of roots of healthy conifers Douglas-fir (Pseudotsuga menziesii) and ponderosa pine (Pinus ponderosa) has been surveyed in the dry forests on the eastern slope of the Cascade Mountains in Washington, USA (Hoff et al. 2004). A total of 27 fungal genera were isolated and identified using a combination of morphological and molecular (ITS region sequences) methods. Fourteen genera were isolated from ponderosa pine, and nine genera from Douglas-fir. Most of the fungi isolated are ascomycetes and zygomycetes, and a few are basidiomycetes. Of these, endophytic fungi Byssochlamys nivea, Umbelopsis spp., and Mucor sp. are the most frequently recovered fungi from ponderosa pine and Douglas-fir. Similarly, a new endophytic fungus Pestalotiopsis hainanensis isolated from healthy branches of Podocarpus macrophyllus in tropical China was identified by a combination of morphological characteristics and ITS and tub2 sequence analyses (Liu et al. 2007). Based on ITS rDNA sequence similarity (95%) to operationally designate species boundaries, a total of 277 fungal species were recovered from 1,403 endophytic strains isolated from common plants in arctic, boreal, temperate, and tropical localities, which represent phylogenetically diverse plant taxa (Arnold and Lutzoni 2007). Similarly, a total of 439 isolates representing 24 morphotaxa were isolated from asymptomatic foliage of loblolly pine (Pinus taeda) in North Carolina, USA. Sequence data from ITS region for 150 isolates revealed 59 distinct ITS genotypes that represent 24 and 37 unique groups based on 90% and 95% sequence similarity, respectively (Arnold et al. 2007). In some studies not only the ITS region but also 18S and 28S rDNA fragments have been employed in the identification of endophytic fungi at various taxonomic levels. Diversity of endophytic fungi isolated from bamboos Phyllostachy and Sasa species were studied based on the analyses of 18S rDNA gene and ITS region sequences, and 71 representative strains were placed into Sordariomycetes and Dothideomycetes. Of these, fungi Xylariales is the dominant group within bamboos and several rDNA gene sequences are not similar to any current sequence in the database and might be novel species or genera (Morakotkarn et al. 2007). Similarly, a total of 47 distinct genotype groups based on 90% ITS sequence similarity were 284 L.-D. Guo obtained from 280 representative strains isolated from healthy photosynthetic tissues of three plant species (Huperzia selago, Picea mariana, and Dryas integrifolia) in northern and southern boreal forests and arctic tundra (Higgins et al. 2007). Further phylogenetic analyses of combined data from 18S and 28S rDNA show that these different genotypic endophytic fungi represent Dothideomycetes, Sordariomycetes, Chaetothyriomycetidae, Leotiomycetes, and Pezizomycetes of Ascomycota. 13.3.2 Identification of Nonsporulating Endophytic Fungi In traditional cultivation-dependent process of endophytic studies, endophytic isolates can be identified only on the basis of morphological characteristics if they sporulate on the media. Despite the development of various methods to promote sporulation, e.g., by growing them on modifications of artificial media and under various incubation conditions (Guo et al. 1998, 2000; Taylor et al. 1999), the number of isolates that do not sporulate ranges from 4.5– 54% of the total isolates (Petrini et al. 1982, Espinosa-Garcia and Langenheim 1990; Johnson and Whitney 1992; Fisher et al. 1993; Guo et al. 2000, 2008; Photita et al. 2001; Cannon and Simmons 2002; Kumaresan and Suryanarayanan 2002; Wang and Guo 2007; Sun et al. 2008). Since conventional classification of fungi relies heavily on reproductive structures, these nonsporulating strains cannot be provided with taxonomic names. In order to appreciate the considerable diversity of these mycelia sterilia in culture, they are generally categorized as “morphotype” on the basis of similar cultural characters (Taylor et al. 1999; Guo et al. 2000, 2003; Arnold et al. 2001; Wang et al. 2005). Arrangement of taxa into different morphotypes, however, does not reflect species phylogeny, because morphotypes are not real taxonomic entities (Lacap et al. 2003; Guo et al. 2000; 2003; Wang et al. 2005). Molecular methods are therefore required for the identification and understanding of the diversity of these endophytic mycelia sterilia. In our survey of endophytic fungi from fronds of Livistona chinensis in Hong Kong, a large number of isolates (16.5% of total isolates) do not sporulate, remaining as mycelia sterilia (Guo et al. 2000). These nonsporulating isolates were grouped into 19 morphotypes on the basis of their cultural morphology. Furthermore, nine morphotypes were identified to genus level (Diaporthe, Mycosphaerella, and Xylaria), five to family level (Pleosporaceae and Clypeosphaeriaceae), and the other five to ordinal level (Xylariales) on the basis of ITS sequence similarity comparisons and phylogenetic analyses. Similarly, in our another study of endophytic fungi of P. tabulaeformis in two distinct climatic sites of Liaoning province of China, a large number of isolates (11% of total isolates) remained as mycelia sterilia (Wang and Guo 2007). These nonsporulating isolates were grouped into 74 morphotypes according to their cultural morphology, and were further divided into 64 taxa on the basis of ITS sequence analyses. Of these morphotypes, five are Basidiomycota and 69 are Ascomycota, and then two morphotypes were identified as Fusarium sporotrichioides and Schizophyllum commune, respectively. 13 Molecular Diversity and Identification of Endophytic Fungi 285 Twenty-two morphotypes were identified to generic level, seven to family (Lophiostomataceae and Valsaceae) level, and four to ordinal (Helotiales and Pezizales) level (Wang et al. 2005). Fifty-nine of morphologically unidentifiable strains isolated from healthy stems and pods of cacao (Theobroma cacao) trees in natural forest ecosystems and agroecosystems in Latin America and West Africa were identified on the basis of the sequence analyses of 28S rDNA (Crozier et al. 2006). The majority of the isolates tested belong to Basidiomycota, particularly to corticoid and polyporoid taxa. Some isolates come from rarely isolated genera, such as Byssomerulius, whilst the most commonly isolated basidiomycetous endophyte is a member of the cosmopolitan genus Coprinellus of Agaricales. In our recent study of endophytic fungi of 20 lichen species in four sites of China, a total of 340 isolates (17.9% of total isolates) did not produce any spores and were divided into 51 morphotypes according to similar cultural characteristics. These morphotypes were placed into 42 taxa, including Atheliales and Agaricales of Basiodiomycota and Coniochaetales, Hypocreales, Pezizales, Pleosporales, Sordariales, and Xylariales of Ascomycota on the basis of ITS sequence analyses (W.C. Li and L.D. Guo, unpublished data). 13.3.3 Identification of White Morphotype Strains In our investigation of endophyte diversity from P. tabulaeformis at Dongling Mountain mixed woodland, the Beijing Forest Ecosystem Research Station of the Chinese Academy of Sciences in China (Guo et al. 2008), the sterile mycelia were divided into different morphotypes on the basis of similar cultural characteristics. Although some sterile isolates having similar cultural characters were grouped into the same morphotype, these isolates might be distantly related taxa. Therefore, an attempt was carried out to establish whether the isolates included in the same morphotype were of the same fungal origin. A total of 18 sterile strains grouped in the white morphotype were selected to evaluate the fungal origins of different isolates using ITS sequence analyses (Guo et al. 2003). Molecular identification showed that five strains belonged to species of Rhytismataceae, and the other 13 strains were identified to Rosellinia, Entoleuca, and Nemania of Xylariaceae (Fig. 13.1). Our results indicate that strains grouped into white morphotype have different fungal origins. 13.3.4 Detection and Identification of Endophytic Fungi Within Plant Tissues Because of the limitations of traditional isolation techniques, it is highly probable that some or even numerous endophytic fungi are never isolated. This may be 286 L.-D. Guo Fig. 13.1 One of 1,720 equally parsimonious trees generated from the ITS (ITS1, 5.8S and ITS2) sequences of 48 taxa showing the relationships of 13 white morphotype strains with reference taxa. The tree rooted with Amphisphaeria umbrina, Discostroma tosta, and Lepteutypa cupressi (Tree length ¼ 1703, Consistency index ¼ 0.437, Homoplasy index ¼ 0.563, Retention index ¼ 0.66, Rescaled consistency index ¼ 0.289). Bootstrap values greater than or equal to 50% (1,000 replicates) are shown at branches. Asterisks indicate the branches that collapse in the strict consensus tree 13 Molecular Diversity and Identification of Endophytic Fungi 287 because some endophytic fungi cannot grow on the artificial media. Most of the endophytic fungi isolated are also usually ascomycetes or their anamorphs. It is not clear whether this is because isolation techniques preclude other fungi, or whether only ascomycetes or their anamorphs constitute the endophytic fungal community. In order to overcome the potential technical bias, molecular techniques have been employed in the detection and identification of endophytic fungi including culturable and nonculturable fungi from the hosts. The molecular study generally includes five steps: (1) The total genomic DNA (including fungi and plants) is extracted from sterile plant tissues; (2) DNA fragments (e.g., ITS, 28S and 18S rDNA) are amplified from total DNA with fungal primers; (3) Polymerase chain reaction (PCR) products (bands) are separated by DGGE or are cloned into plasmids (e.g., pGEM-T vector); (4) Different single clones are screened using DNA fingerprinting techniques (e.g., RFLP and SSCP) and different DGGE bands are excised; (5) Representative clones and DGGE bands are sequenced and theoretically identified into various taxonomic levels on the basis of phylogenetic analysis and sequence similarity comparison. Endophytic fungal community of Marram grass (Ammophila arenaria) roots were analyzed using DGGE with subsequent cloning and sequencing to identify the fungi by amplification of partial 18S rDNA gene (Kowalchuk et al. 1997). Some ITS fragments amplified from Picea foliages were identified as endophytic fungi isolated from the same plant tissues using a cultivation-dependent method (Camacho et al. 1997). In our study, fungal ITS regions were amplified directly from total genomic DNA extracted from fronds of Livistona chinensis. A total of five different cloned sequences of fungi were obtained; of these four cloned sequences were identified as Glomerella (anamorph Colletotrichum), Mycosphaerella (anamorph Cladosporium), and Herpotrichiellaceae of Ascomycetes, and the other one cloned sequence belonged to Basidiomycetes which is not found using traditional cultivation-dependent method (Guo et al. 2001). The variation in endophytic fungal diversity closely associated with roots, stems, and leaves of common reed (Phragmites australis) from two dry and two flooded sites at Lake Constance in Germany were investigated on the basis of ITS sequence analysis (Wirsel et al. 2001). Most isolates were Ascomycetes, and some were Basidiomycetes. The result indicates that there are differences in distribution of endophytic fungi between dry and flooded sites. Similarly, the differences in the composition and relative abundance of endophytic fungi colonizing the roots of T4-lysozyme producing potatoes and the parental line were detected by amplified 18S rRNA gene fragments from total community DNA extracted from roots of potato Solanum tuberosum (Götz et al. 2006). In the detection of endophytic fungi of Heterosmilax japonica tissues, a broad spectrum of fungal ITS sequences was directly amplified from genomic DNA extracted from host tissues (Gao et al. 2005). Of these fungal sequences some were identified as Aureobasidium, Botryosphaeria, Cladosporium, Glomerella, Mycosphaerella, Phomopsis, and Guignardia, the others (e.g., YJ4-61, YJ4-9 and YJ4-70) were significantly similar to some uncultured environmental samples and were not specifically affiliated with any currently documented fungal sequences in 288 L.-D. Guo the NBCI GenBank database. Endophytic fungal rDNA fragments (28S and ITS) were amplified from surface sterilized needles from 12 Pinus taeda trees in North Carolina, USA (Arnold et al. 2007). Phylogenetic analyses of 28S rDNA indicate that cloned endophytic fungi are distributed across multiple lineages of Ascomycota and Basidiomycota. Further identification of cloned endophytic fungi based on ITS sequence analyses shows that there are at least four unique fungal species within Basidiomycota and at least nine fungal species within Ascomycota. Ascomycetous endophytic fungi are primarily Dothideomycetes and Leotiomycetes, which are commonly isolated from P. taeda using traditional cultural methods, but no Sordariomycetes were recovered from cloned endophytic sequences, despite the prevalence of this lineage among cultural endophytes. The results of some previous studies show that the diversity of endophytic fungi detected with molecular methods differs from that found using traditional cultivation-dependent methods. Although molecular techniques insight into diversity of endophytic fungi, there is disadvantage in the identification of endophyte morphotypes based on ITS sequence analysis. There is no criterion to delimit species boundary of ITS sequence divergence. This is because, although some endophytic fungi have high similarity in the ITS sequences and cluster together with high bootstrap support with reference taxa, there is still insufficient information at present to determine whether the terminal clades include one or more species in the phylogenetic analysis. For most of the taxa included in the DNA sequence analyses, the level of interspecific and intraspecific variations is still variable. Different levels of variations have been reported in the different taxa in previous studies. A relatively low substitution rate was reported in ITS sequences of several Armillaria species (0.5%) from the Northern Hemisphere (Anderson and Stasovski 1992) and among Sclerotium species (Carbone and Kohn 1993). On the contrary, there is a relatively low level of homology (76.1%) between weakly virulent and highly virulent isolates of Leptosphaeria maculans, while the ITS sequences differ in only four nucleotide positions within the highly virulent isolates and in two nucleotide positions within the weakly virulent isolates (Morales et al. 1993). Similarly, there is great divergence among three ITS types of Fusarium sambucinum (4.6–15%), while the divergence is extremely low (0–2.3%) within each type (O’Donnell 1992). Significantly, Arnold et al. (2007) have constructed well-supported phylogenies based on a ca. 600 bp of the 28S rDNA for 72 Ascomycota and Basidiomycota, 145 cultured endophytic fungi, and 33 environmental PCR samples. The result shows that ITS genotype groups based on 90% sequence similarity are concordant with 28S rDNA-delimited species. However, at present there appears to be absence of definite criteria for interspecific and intraspecific level of nucleotide divergence in ITS region sequences of fungi. There are some limitations of the detection and identification of endophytic fungi directly from within plant tissues using molecular techniques. Firstly, as only sparse hyphae may exist within the plant tissues, some fungal DNA may be lost during the DNA extraction process, thus only a minor fraction of fungal DNA is included in the total DNA extracted from plant tissues. Secondly, there are 13 Molecular Diversity and Identification of Endophytic Fungi 289 inhibitors that may interfere with the PCR amplification in the DNA solution. Thirdly, the universal primers may not completely match with some fungal template DNA. In addition, it is important to take into consideration that surface sterilization may not have denaturized the DNA of epiphytes, although sodium hypochlorite is relatively effective for this purpose. Therefore, it is likely to be difficult to amplify all endophytic fungal DNA fragments from the total DNA samples. Another limitation is the limited number of sequences, i.e., less than 1%, of the estimated 1.5 million fungal species presented in NBCI GenBank and EMBL database, although there is daily increase in fungal DNA sequences in public databases (Vilgalys 2003). In addition, misidentifications of named published sequences, of which ca. 20% of the named sequences may be attributed to incorrectly named organisms, may represent another problem restricting the feasibility of sequencebased identification of endophytic fungi (Vilgalys 2003; Hawksworth 2004). 13.4 Conclusions Molecular fingerprinting techniques are powerful tools in the detection of population genetic structure and diversity of endophytic fungi. Further development of these markers to allow detection of endophytes in planta will considerably enhance their value, and will permit the sensitive detection of endophyte incidence in plant populations. Molecular sequencing techniques offer an effective method for the identification of endophytic fungi, particularly for nonsporulating isolates, and for the detection of the viable but nonculturable fungi by directly amplified rDNA fragments from plant tissues. PCR-based molecular techniques are conventional PCR employed in the detection and identification of endophytic fungi in previous studies. These conventional PCR can identify endophytic fungi specifically, but it cannot be used to quantify endophytic fungal biomass within plant tissues. However, real-time PCR can detect small quantities of DNA in environmental samples and has been successfully used to determine the population density of some fungal species such as Pyrenophora sp. (Bates et al. 2001), Plectosphaerella cucumerina and Paecilomyces lilacinus (Atkins et al. 2003, 2004), and Hirsutella rhossiliensis (Zhang et al. 2006). DNA barcoding systems employ a short, effective, standardized gene region to identify species (Hebert et al. 2003; Blaxter 2003; Savolainen et al. 2005; Seifert et al. 2007; Craig et al. 2008). To date, this technique has been extensively used in the animal kingdom with a 648-bp region of the CO1 gene (Smith et al. 2005; Ward et al. 2005; Hajibabaei et al. 2006). DNA barcodes were first employed in the identification of fungi Penicillium species using CO1 gene (Seifert et al. 2007) and ectomycorrhizal fungi in a Tasmanian wet sclerophyll forest by ITS regions (Tedersoo et al. 2008). With the improvement of molecular techniques, e.g., DNA fingerprinting, DNA sequencing, real-time PCR, and DNA barcoding, they will become routine, accurate, rapid, and sensitive techniques in the detection, identification, and quantification of endophytic fungal diversity in future. 290 L.-D. Guo References Adachi Y, Watanabe H, Tababe K, Doke N, Nishimura S, Tsuge T (1993) Nuclear ribosomal DNA as a probe for genetic variability in the Japanese pear pathotype of Alternaria alternata. Appl Environ Microbiol 59:3197–3205 Anderson IC, Cairney JWG (2004) Diversity and ecology of soil fungal communities: increased understanding through the application of molecular techniques. Environ Microbiol 6:769–779. doi:10.1111/j.1462-2920.2004.00675.x Anderson IC, Campbell CD, Prosser JI (2003) Diversity of fungi in organic soils under a moorland – Scots pine (Pinus sylvestris L.) gradient. Environ Microbiol 5:1121–1132. doi:10.1046/ j.1462-2920.2003.00522.x Anderson JB, Stasovski E (1992) Molecular phylogeny of northern hemisphere species of Armillaria. Mycologia 84:505–516 Aradhya MK, Chan HM, Parfitt DE (2001) Genetic variability in the pistachio late blight fungus, Alternaria alternata. Mycol Res 105:300–306. doi:10.1017/S0953756201003677 Arnold AE, Lutzoni F (2007) Diversity and host range of foliar fungal endophytes: are tropical leaves biodiversity hotspots. Ecology 88:541–549. doi:10.1890/05-1459 Arnold AE, Henk DA, Eells RL, Lutzoni F, Vilgalys R (2007) Diversity and phylogenetic affinities of foliar fungal endophytes in loblolly pine inferred by culturing and environmental PCR. Mycologia 99:185–206. doi:10.3852/mycologia.99.2.185 Arnold AE, Maynard Z, Gilbert GS (2001) Fungal endophytes in dicotyledonous neotropical trees: patterns of abundance and diversity. Mycol Res 105:1502–1507. doi:10.1017/ S0953756201004956 Atkins SD, Clark IM, Pande S, Hirsch PR, Kerry BR (2004) The use of real-time PCR and speciesspecific primers for the identification and monitoring of Paecilomyces lilacinus. FEMS Microbiol Ecol 51:257–264. doi:10.1016/j.femsec.2004.09.002 Atkins SD, Clark IM, Sosnowska D, Hirsch PR, Kerry BR (2003) Detection and quantification of Plectosphaerella cucumerina, a potential biological control agent of potato cyst nematodes, by using conventional PCR, real-time PCR, selective media and baiting. Appl Environ Microbiol 69:4788–4793. doi:10.1128/AEM.69.8.4788-4793.2003 Bates JA, Taylor EJA, Kenyon DM, Thomas JE (2001) The application of real-time PCR to the identification, detection and quantification of Pyrenophora species in barley seed. Mol Plant Pathol 2:49–57. doi:10.1111/j.1364-3703.2001.00049.x Beck JJ, Ligon JM (1995) Polymerase chain reaction assays for the detection of Stagonospora nodorum and Septoria tritici in wheat. Phytopathology 85:319–324. doi:10.1094/Phyto-85-319 Bills GF (1996) Isolation and analysis of endophytic fungal communities from wood plants. In: Redlin SC, Carris LM (eds) Endophytic fungi in grasses and woody plants: systematics, ecology, and evolution. APS Press, St. Paul, Minnesota, pp 31–65 Blaxter M (2003) Counting angels with DNA. Nature 421:122–124. doi:10.1038/421122a Bock CH, Thrall PH, Brubaker CL, Burdon JJ (2002) Detection of genetic variation in Alternaria brassicicola using AFLP fingerprinting. Mycol Res 106:428–434. doi:10.1017/ S0953756202005762 Burgess T, Wingfield MJ, Wingfield BW (2001) Simple sequence repeat markers distinguish among morphotypes of Sphaeropsis sapinea. Appl Environ Microbiol 67:354–362. doi:10.1128/AEM.67.1.354-362.2001 Camacho FJ, Gernandt DS, Liston A, Stone JK, Klein AS (1997) Endophytic fungal DNA, the source of contamination in spruce needle DNA. Mol Ecol 6:983–987. doi:10.1046/j.1365294X.1997.00266.x Cannon PF, Simmons CM (2002) Diversity and host preference of leaf endophytic fungi in the Iwokrama Forest Reserve, Guyana. Mycologia 94:210–220 Carbone I, Kohn LM (1993) Ribosomal DNA sequence divergence within internal transcribed spacer 1 of the Sclerotiniaceae. Mycologia 85:415–427 13 Molecular Diversity and Identification of Endophytic Fungi 291 Carroll GC (1986) The biology of endophytism in plants with particular reference to woody plants. In: Fokkema NJ, van den Heuvel J (eds) Microbiology of the phyllosphere. Cambridge University Press, Cambridge, U.K., pp 205–222 Chambers SM, Sawyer NA, Cairney JWG (1999) Molecular identification of co-occurring Cortinarius and Dermocybe species from southeastern Australian sclerophyll forests. Mycorrhiza 9:85–90. doi:10.1007/s005720050004 Chelius MK, Triplett EW (1999) Rapid detection of arbuscular mycorrhizae in roots and soil of an intensively managed turfgrass system by PCR amplification of small subunit rDNA. Mycorrhiza 9:61–64. doi:10.1007/s005720050264 Clapp JP, Young JPW, Merryweather JW, Fitter AH (1995) Diversity of fungal symbionts arbuscular mycorrhizas from a natural community. New phytol 130:259–265. doi:10.1111/ j.1469-8137.1995.tb03047.x Countway PD, Gast RJ, Savai P, Caron DA (2005) Protistan diversity estimates based on 18S rDNA from seawater incubations in the Western North Atlantic. J Eukaryot Microbiol 52:95–106. doi:10.1111/j.1550-7408.2005.05202006.x Craig DW, Pearson JV, Szelinger S, Sekar A, Redman M, Corneveaux JJ, Pawlowski TL, Laub T, Nunn G, Stephan DA, Homer N, Huentelman MJ (2008) Identification of genetic variants using bar-coded multiplexed sequencing. Nat Methods 5:887–893. doi:10.1038/nmeth.1251 Crozier J, Thomas SE, Aime MC, Evans HC, Holmes KA (2006) Molecular characterization of fungal endophytic morphospecies isolated from stems and pods of Theobroma cacao. Plant Pathol 55:783–791. doi:10.1111/j.1365-3059.2006.01446.x Dar SA, Kuenen JG, Muyzer G (2005) Nested PCR-denaturing gradient gel electrophoresis approach to determine the diversity of sulfate-reducing bacteria in complex microbial communities. Appl Environ Microbiol 71:2325–2330. doi:10.1128/AEM.71.5.2325-2330.2005 De Bary A (1866) Morphologie und Physiologie der Pilze, Flechten, und Myxomyceten. Hofmeister’s handbook of physiological botany, Vol. 2. Leipzig De Jong EZ, Guthridge KM, Spangenberg GC, Forster JW (2003) Development and characterization of EST-derived simple sequence repeat (SSR) markers for pasture grass endophytes. Genome 46:277–290. doi:10.1139/g03-001 Dı́ez B, Pedrós-Alió C, Marsh TL, Massana R (2001) Application of denaturing gradient gel electrophoresis (DGGE) to study the diversity of marine Picoeukaryotic assemblages and comparison of DGGE with other molecular techniques. Appl Environ Microbiol 67:2942–2951. doi:10.1128/AEM.67.7.2942-2951.2001 Duong LM, Jeewon R, Lumyong S, Hyde KD (2006) DGGE coupled with ribosomal DNA gene phylogenies reveal uncharacterized fungal phylotypes. Fungal Divers 23:121–138 Espinosa-Garcia FJ, Langenheim JH (1990) The leaf fungal endophytic community of a coastal redwood population diversity and spatial patterns. New Phytol 116:89–97. doi:10.1111/j.14698137.1990.tb00513.x Fisher PJ, Petrini O, Sutton BC (1993) A comparative study of fungal endophytes in leaves, xylem and bark of Eucalyptus nitens in Australia and England. Sydowia 45:338–345 Gao XX, Zhou H, Xu DY, Yu CH, Chen YQ, Qu LH (2005) High diversity of endophytic fungi from the pharmaceutical plant, Heterosmilax japonica Kunth revealed by cultivation-independent approach. FEMS Microbiol Lett 249:255–266. doi:10.1016/j.femsle.2005.06.017 Garcı́a RA, Zapater JMM, Criado BG, Zabalgogeazcoa I (2002) Genetic structure of natural populations of the grass endophyte Epichloe¨ festucae in semiarid grasslands. Mol Ecol 11:355–364. doi:10.1046/j.0962-1083.2001.01456.x Gardes M, White TJ, Fortin JA, Bruns TD, Taylor JW (1991) Identification of indigenous and introduced symbiotic fungi in ectomycorrhizae by amplification of nuclear and mitochondrial ribosomal DNA. Can J Bot 69:180–190. doi:10.1139/b91-026 Götz M, Nirenberg H, Krause S, Wolters H, Draeger S, Buchner A, Lottmann J, Berg G, Smalla K (2006) Fungal endophytes in potato roots studied by traditional isolation and cultivationindependent DNA-based methods. FEMS Microbiol Ecol 58:404–413. doi:10.1111/j.15746941.2006.00169.x 292 L.-D. Guo Groppe K, Boller T (1997) PCR assay based on a microsatellite-containing locus for detection and quantification of Epichloe¨ endophytes in grass tissue. Appl Environ Microbiol 63:1543–1550 Groppe K, Sanders I, Wiemken A, Boller T (1995) A microsatellite marker for studying the ecology and diversity of fungal endophytes (Epichloe¨ spp.) in grasses. Appl Environ Microbiol 61:3943–3949 Grünig CR, Brunner PC, Duò A, Sieber TN (2007) Suitability of methods for species recognition in the Phialocephala fortinii-Acephala applanata species complex using DNA analysis. Fungal Genet Biol 44:773–788. doi:10.1016/j.fgb.2006.12.008 Grünig CR, Duò A, Sieber TN (2006) Population genetic analysis of Phialocephala fortinii s.l. and Acephala applanata in two undisturbed forests in Switzerland and evidence for new cryptic species. Fungal Genet Biol 43:410–421. doi:10.1016/j.fgb.2006.01.007 Grünig CR, Duò A, Sieber TN, Holdenrieder O (2008) Assignment of species rank to six reproductively isolated cryptic species of the Phialocephala fortinii s.l.-Acephala applanata species complex. Mycologia 100:47–67. doi:10.3852/mycologia.100.1.47 Grünig CR, Linde CC, Sieber TN, Rogers SO (2003) Development of single-copy RFLP markers for population genetic studies of Phialocephala fortinii and closely related taxa. Mycol Res 107:1332–1341. doi:10.1017/S0953756203008669 Grünig CR, McDonald BA, Sieber TN, Rogers SO, Holdenrieder O (2004) Evidence for subdivision of the root-endophyte Phialocephala fortinii into cryptic species and recombination within species. Fungal Genet Biol 41:676–687. doi:10.1016/j.fgb.2004.03.004 Grünig CR, Sieber TN, Holdenrieder O (2001) Characterisation of dark septate endophytic fungi (DSE) using inter-simple-sequence-repeat-anchored polymerase chain reaction (ISSR-PCR) amplification. Mycol Res 105:24–32. doi:10.1017/S0953756200003658 Grünig CR, Sieber TN, Rogers SO, Holdenrieder O (2002) Spatial distribution of dark septate endophytes in a confined forest plot. Mycol Res 106:832–840. doi:10.1017/S0953756202005968 Guo LD (2001) Advances of researches on endophytic fungi. Mycosystema 20:148–152 Guo LD, Huang GR, Wang Y (2008) Seasonal and tissue age influences on endophytic fungi of Pinus tabulaeformis (Pinaceae) in Dongling Mountain, Beijing. J Integr Plant Biol 50:997–1003. doi:10.1111/j.1744-7909.2008.00394.x Guo LD, Huang GR, Wang Y, He WH, Zheng WH, Hyde KD (2003) Molecular identification of white morphotype strains of endophytic fungi from Pinus tabulaeformis. Mycol Res 107:680–688. doi:10.1017/S0953756203007834 Guo LD, Hyde KD, Liew ECY (1998) A method to promote sporulation in palm endophytic fungi. Fungal Divers 1:109–113 Guo LD, Hyde KD, Liew ECY (2000) Identification of endophytic fungi from Livistona chinensis (Palmae) using morphological and molecular techniques. New Phytol 147:617–630. doi:10.1046/j.1469-8137.2000.00716.x Guo LD, Hyde KD, Liew ECY (2001) Detection and identification of endophytic fungi within frond tissues of Livistona chinensis based on rDNA sequence. Mol Phylogenet Evol 20:1–13. doi:10.1006/mpev.2001.0942 Guo LD, Xu L, Zheng WH, Hyde KD (2004) Genetic variation of Alternaria alternata, an endophytic fungus isolated from Pinus tabulaeformis as determined by random amplified microsatellites (RAMS). Fungal Divers 16:53–65 Hajibabaei M, Janzen DH, Burns JM, Hallwachs W, Hebert PDN (2006) DNA barcodes distinguish species of tropical Lepidoptera. PNAS 103:968–971. doi:10.1073/pnas.0510466103 Hämmerli UA, Brändle UE, Petrini O, McDermott JM (1992) Differentiation of isolates of Discula umbrinella (teleomorph: Apiognomonia errabunda) from beech, chestnut and oak using RAPD markers. Mol Plant Microbe 5:479–483. doi:10.1094/MPMI-5-479 Hantula J, Müller MM (1997) Variation within Cremmeniella abietina in Finland and other countries as determined by random amplified microsatellites (RAMS). Mycol Res 101:169–175. doi:10.1017/S0953756296002225 13 Molecular Diversity and Identification of Endophytic Fungi 293 Hantula J, Dusabenyagasani M, Hamelin RC (1996) Random amplified microsatellites (RAMS) – a novel method for characterizing genetic variation within fungi. Eur J Forest Pathol 26:159–166. doi:10.1111/j.1439-0329.1996.tb00720.x Hawksworth DL (2004) ‘Misidentifications’ in fungal DNA sequence databanks. New Phytol 161:13–15. doi:10.1111/j.1469-8137.2004.00958.x Hawksworth DL (1991) The fungal dimension of biodiversity: magnitude, significance, and conservation. Mycol Res 95:641–655. doi:10.1016/S0953-7562(09)80810-1 Hebert PDN, Cywinska A, Ball SL, de Waard JR (2003) Biological identifications through DNA barcodes. Proc R Soc Lond B 270:313–321. doi:10.1098/rspb.2002.2218 Higgins KL, Arnold AE, Miadlikowska J, Sarvate SD, Lutzoni F (2007) Phylogenetic relationships, host affinity, and geographic structure of boreal and arctic endophytes from three major plant lineages. Mol Phylogenet Evol 42:543–555. doi:10.1016/j.ympev.2006.07.012 Hoff JA, Klopfenstein NB, McDonald GI, Tonn JR, Kim MS, Zambino PJ, Hessburg PF, Rogers JD, Peever TL, Carris LM (2004) Fungal endophytes in woody roots of Douglas-fir (Pseudotsuga menziesii) and ponderosa pine (Pinus ponderosa). Forest Pathol 34:255–271. doi:10.1111/j.1439-0329.2004.00367.x Huai WX, Guo LD, He W (2003) Genetic diversity of an ectomycorrhizal fungus Tricholoma terreum in Larix principis-rupprechtii stand assessed using RAPDs. Mycorrhiza 13:265–270. doi:10.1007/s00572-003-0227-8 Jansa J, Mozafar A, Kuhn G, Anken T, Ruh R, Sanders IR, Frossard E (2003) Soil tillage affects the community structure of mycorrhizal fungi in maize roots. Ecol Appl 13:1164–1176. doi:10.1890/1051-0761(2003) 13[1164:STATCS]2.0.CO;2 Jeewon R, Hyde KD (2006) Diversity and detection of fungi from environmental samples: traditional versus molecular approaches. In: Varma A, Oelmuller R (eds) Advanced techniques in soil microbiology. Soil Biology Series. Springer, Berlin, pp 1–15 Johnson JA, Whitney NJ (1992) Isolation of fungal endophytes from black spruce (Picea mariana) dormant buds and needles from New Brunswick. Can J Bot 70:1754–1757. doi:10.1139/b92-217 Jumpponen A (1999) Spatial distribution of discrete RAPD phenotypes of a root endophytic fungus, Phialocephala fortinii, at a primary successional site on a glacier forefront. New Phytol 141:333–344. doi:10.1046/j.1469-8137.1999.00344.x Klamer M, Roberts MS, Levine LH, Drake BG, Garland JL (2002) Influence of elevated CO2 on the fungal community in a coastal scrub oak forest soil investigated with terminal restriction fragment length polymorphism analysis. Appl Environ Microbiol 68:4370–4376. doi:10.1128/ AEM.68.9.4370-4376.2002 Kowalchuk G, Gerards S, Woldendorp JW (1997) Detection and characterization of fungal infections of Ammophila arenaria (Marram grass) roots by denaturing gradient gel electrophoresis of specifically amplified 18S rDNA. Appl Environ Microbiol 63:3858–3865 Kumaresan V, Suryanarayanan TS (2002) Endophytic assemblages in young, mature and senescent leaves of Rhizophora apiculata: evidence for the role of endophytes in mangrove litter degradation. Fungal Divers 9:91–91 Lacap DC, Hyde KD, Liew ECY (2003) An evaluation of the fungal ‘morphotype’ concept based on ribosomal DNA sequences. Fungal Divers 12:53–66 Levinson G, Gutman GA (1987) Slipped-strand mispairing: a major mechanism for DNA sequence evolution. Mol Biol Evol 4:203–221 Li WC, Zhou J, Guo SY, Guo LD (2007) Endophytic fungi associated with lichens in Baihua mountain of Beijing, China. Fungal Divers 25:69–80 Liang Y, Guo LD, Ma KP (2004) Genetic structure of a population of the ectomycorrhizal fungus Russula vinosa in subtropical woodlands in southwest China. Mycorrhiza 14:235–240. doi:10.1007/s00572-003-0260-7 Liang Y, Guo LD, Ma KP (2005) Population genetic structure of an ectomycorrhizal fungus Amanita manginiana in a subtropical forest over 2 years. Mycorrhiza 15:137–142. doi:10.1007/s00572-004-0311-8 294 L.-D. Guo Liu AR, Xu T, Guo LD (2007) Molecular and morphological description of Pestalotiopsis hainanensis sp. nov., a new endophyte from a tropical region of China. Fungal Divers 24:23–36 Ma LJ, Catramis CM, Rogers SO, Starmer WT (1997) Isolation and characterisation of fungi entrapped in glacial ice. Inoculum 48:23–24 May LA, Smiley B, Schmidt MG (2001) Comparative denaturing gradient gel electrophoresis analysis of fungal communities associated with whole plant corn silage. Can J Microbiol 47:829–841. doi:10.1139/cjm-47-9-829 McCutcheon TL, Carroll GC (1993) Genotypic diversity in populations of a fungal endophyte from Douglas fir. Mycologia 85:180–186 Mills PR, Sreenivasaprasad S, Brown AE (1992) Detection and differentiation of Colletotrichum gloeosporioides isolates using PCR. FEMS Microbiol Lett 98:137–144. doi:10.1111/j.15746968.1992.tb05503.x Mohali S, Burgess TI, Wingfield MJ (2005) Diversity and host association of the tropical tree endophyte Lasiodiplodia theobromae revealed using simple sequence repeat markers. Forest Pathol 35:385–396. doi:10.1111/j.1439-0329.2005.00418.x Morakotkarn D, Kawasaki H, Seki T (2007) Molecular diversity of bamboo-associated fungi isolated from Japan. FEMS Microbiol Lett 266:10–19. doi:10.1111/j.1574-6968.2006.00489.x Morales VM, Pelcher LE, Taylor JL (1993) Comparison of the 5.8 S rDNA and internal transcribed spacer sequences of isolates of Leptosphaeria maculans from different pathogenicity groups. Curr Genet 23:490–495. doi:10.1007/BF00312640 Moukhamedov R, Hu X, Nazar RN, Robb J (1994) Use of polymerase chain reaction-amplified ribosomal intergenic sequences for the diagnosis of Verticillium tricorpus. Phytopathology 84:256–259. doi:10.1094/Phyto-84-256 Nikolcheva LG, Cockshutt AM, Bärlocher F (2003) Determining diversity of freshwater fungi on decaying leaves: comparison of traditional and molecular approaches. Appl Environ Microbiol 69:2548–2554. doi:10.1128/AEM.69.5.2548-2554.2003 O’Donnell K (1992) Ribosomal DNA internal transcribed spacers are highly divergent in the phytopathogenic ascomycete Fusarium sambucinum (Gibberella pulicaris). Curr Genet 22:213–220 Pandey AK, Reddy MS, Suryanarayanan TS (2003) ITS-RFLP and ITS sequence analysis of a foliar endophytic Phyllosticta from different tropical trees. Mycol Res 107:439–444. doi:10.1017/S0953756203007494 Peever TL, Ibañez A, Akimitsu K, Timmer LW (2002) Worldwide phylogeography of the citrus brown spot pathogen, Alternaria alternata. Phytopathology 92:794–802. doi:10.1094/ PHYTO.2002.92.7.794 Petrini O (1991) Fungal endophytes of tree leaves. In: Andrews JH, Hirano SS (eds) Microbial ecology of leaves. Springer-Verlag, New York, pp 179–197 Petrini O, Stone J, Carroll FE (1982) Endophytic fungi in evergreen shrubs in western Oregon: a preliminary study. Can J Bot 60:789–796. doi:10.1139/b82-102 Photita W, Lumyong S, Lumyong P, Hyde KD (2001) Endophytic fungi of wild banana (Musa acuminata) at doi Suthep Pui National Park, Thailand. Mycol Res 105:1508–1513. doi:10.1017/S0953756201004968 Ranjard L, Poly F, Nazaret S (2000) Monitoring complex bacterial communities using culture-independent molecular techniques: application to soil environment. Res Microbiol 151:167–177. doi:10.1016/S0923-2508(00)00136-4 Roberts RG, Reymond ST, Andersen B (2000) RAPD fragment pattern analysis and morphological segregation of small-spored Alternaria species and species groups. Mycol Res 104:151–160. doi:10.1017/S0953756299001690 Rodrigues KF, Sieber TN, Grunig CR, Holdenrieder O (2004) Characterization of Guignardia mangiferae isolated from tropical plants based on morphology, ISSR-PCR amplifications and ITS1–5.8 S-ITS2 sequences. Mycol Res 108:45–52. doi:10.1017/S0953756203008840 13 Molecular Diversity and Identification of Endophytic Fungi 295 Rollo F, Sassaroli S, Ubaldi M (1995) Molecular phylogeny of the fungi of the Iceman’s grass clothing. Curr Genet 28:289–297. doi:10.1007/BF00309789 Savolainen V, Cowan RS, Vogler AP, Roderick GK, Lane R (2005) Towards writing the encyclopaedia of life: an introduction to DNA barcoding. Phil Trans R Soc B 360:1805– 1811. doi:10.1098/rstb.2005.1730 Schesser K, Luder A, Henson JM (1991) Use of polymerase chain reaction to detect the take-all fungus, Gaeumannomyces graminis, in infected wheat plants. Appl Environ Microbiol 57:553–556 Schmalenberger A, Tebbe CC (2003) Bacterial diversity in maize rhizospheres: conclusions on the use of genetic profiles based on PCR-amplified partial small subunit rRNA genes in ecological studies. Mol Ecol 12:251–262. doi:10.1046/j.1365-294X.2003.01716.x Schulz B, Boyle C (2005) The endophytic continuum. Mycol Res 109:661–686. doi:10.1017/ S095375620500273X Schulz B, Boyle C (2006) What are endophytes. In: Schul ZB, Boyle C, Sieber T (eds) Microbial root endophytes. Springer, Berlin, pp 1–13 Seifert KA, Samson RA, de Waard JR, Houbraken J, Lévesque CA, Moncalvo JM, Louis-Seize G, Hebert PDN (2007) Prospects for fungus identification using CO1 DNA barcodes, with Penicillium as a test case. PNAS 104:3901–3906. doi:10.1073/pnas.0611691104 Simon L, Lévesque RC, Lalonde M (1993) Identification of endomycorrhizal fungi colonizing roots by fluorescent single-strand conformation polymorphism-polymerase chain reaction. Appl Environ Microbiol 59:4211–4215 Sinclair JB, Cerkauskas RF (1996) Latent infection vs. endophytic colonization by fungi. In: Redlin SC, Carris LM (eds) Endophytic fungi in grasses and woody plants: systematics, ecology, and evolution. Minnesota, APS Press, St. Paul, pp 3–29 Smit E, Leeflang P, Glandorf B, van Elsas JD, Wernars K (1999) Analysis of fungal diversity in the wheat rhizosphere by sequencing of cloned PCR-amplified genes encoding 18S rRNA and temperature gradient gel electrophoresis. Appl Environ Microbiol 65:2614–2621 Smith MA, Fisher BL, Hebert PDN (2005) DNA barcoding for effective biodiversity assessment of a hyperdiverse arthropod group: the ants of Madagascar. Phil Trans R Soc B 360:1825–1834. doi:10.1098/rstb.2005.1714 Stephen JR, Kowalchuk GA, Bruns MAV, McCaig AE, Phillips CJ, Embley TM, Prosser JI (1998) Analysis of b-subgroup proteobacterial ammonia oxidizer populations in soil by denaturing gradient gel electrophoresis and hierarchical phylogenetic probing. Appl Environ Microbiol 64:2958–2965 Stone JK, Bacon CW, White JF (2000) An overview of endophytic microbes: endophytism defined. In: Bacon CW, White JF (eds) Microbial endophytes. Marcel Dekker, New York, pp 3–30 Sun JQ, Guo LD, Zang W, Ping WX, Chi DF (2008) Diversity and ecological distribution of endophytic fungi associated with medicinal plants. Sci China Ser C Life Sci 51:751–759 Sun X, Guo LD (2007) Endophytic fungi VI. Ciliophora quercus sp. nov. from China. Nova Hedwigia 85:403–406. doi:10.1127/0029-5035/2007/0085-0403 Taylor JE, Hyde KD, Jones EBG (1999) Endophytic fungi associated with the temperate palm, Trachycarpus fortunei, within and outside its natural geographic range. New Phytol 142:335–346. doi:10.1046/j.1469-8137.1999.00391.x Tedersoo L, Jairus T, Horton BM, Abarenkov K, Suvi T, Saar I, Kõljalg U (2008) Strong host preference of ectomycorrhizal fungi in a Tasmanian wet sclerophyll forest as revealed by DNA barcoding and taxon-specific primers. New Phytol 180:479–490. doi:10.1111/j.14698137.2008.02561.x Tooley PW, O’Neill NR, Goley ED, Carras MM (2000) Assessment of diversity in Claviceps africana and other Claviceps species by RAM and AFLP analyses. Phytopathology 90:1126–1130. doi:10.1094/PHYTO.2000.90.10.1126 Vainio EJ, Hantula J (2000) Direct analysis of wood-inhabiting fungi using denaturing gradient gel electrophoresis. Mycol Res 104:927–930. doi:10.1017/S0953756200002471 296 L.-D. Guo van Elsas JD, Duarte GF, Keijzer-Wolters A, Smit E (2000) Analysis of the dynamics of fungal communities in soil via fungal-specific PCR of soil DNA followed by denaturing gradient gel electrophoresis. J Microbiol Methods 43:133–151. doi:10.1016/S0167-7012(00)00212-8 Vilgalys R (2003) Taxonomic misidentification in public DNA database. New Phytol 160:4–5. doi:10.1046/j.1469-8137.2003.00894.x Wang Y, Guo LD (2007) A comparative study of endophytic fungi in needles, bark, and xylem of Pinus tabulaeformis. Can J Bot 85:911–917. doi:10.1139/B07-084 Wang Y, Guo LD, Hyde KD (2005) Taxonomic placement of sterile morphotypes of endophytic fungi from Pinus tabulaeformis (Pinaceae) in northeast China based on rDNA sequences. Fungal Divers 20:235–260 Ward RD, Zemlak TS, Innes BH, Last PR, Hebert PDN (2005) DNA barcoding Australia’s fish species. Phil Trans R Soc B 360:1847–1857. doi:10.1098/rstb.2005.1716 Wei JG, Xu T, Guo LD, Liu AR, Zhang Y, Pan XH (2007) Endophytic Pestalotiopsis species associated with plants of Podocarpaceae, Theaceae and Taxaceae in southern China. Fungal Divers 24:55–74 Wilson D (1995) Endophyte: the evolution of a term, and clarification of its use and definition. Oikos 73:274–276 Wirsel SG, Liebinger W, Ernst M, Mendgen K (2001) Genetic diversity of fungi closely associated with common reed. New Phytol 149:589–598. doi:10.1046/j.1469-8137.2001.00038.x Zhang LM, Liu XZ, Zhu SF, Chen SY (2006) Detection of the nematophagous fungus Hirsutella rhossiliensis in soil by real-time PCR and parasitism bioassay. Biol Control 36:316–323. doi:10.1016/j.biocontrol.2005.08.002 Zhang W, Wendel JF, Clark LG (1997) Bamboozled again! Inadvertent isolation of fungal rDNA sequences from bamboos (Poaceae: Bambusoideae). Mol Phylogenet Evol 8:205–217. doi:10.1006/mpev.1997.0422 Zheng RY, Jiang H (1995) Rhizomucor endophyticus sp. nov., an endophytic zygomycetes from higher plants. Mycotaxon 56:455–466 Zietkiewicz E, Rafalski A, Labuda D (1994) Genome fingerprinting by simple sequence repeat (SSR)-anchored polymerase chain reaction amplification. Genomics 20:176–183. doi:10.1006/ geno.1994.1151 Chapter 14 Molecular Identification of Anaerobic Rumen Fungi Martin Eckart, Katerina Fliegerová, Kerstin Hoffmann, and Kerstin Voigt Abstract Anaerobic fungi are phylogenetically unique and form a separate group, the Neocallimastigomycota, among the chitinous fungi. Until now six genera are described within that phylum, namely the monocentric genera Neocallimastix, Caecomyces and Piromyces as well as the polycentric genera Anaeromyces, Cyllamyces and Orpinomyces. This chapter gives a brief survey of the fascinating world of anaerobic rumen fungi, their phylogeny, and identification. The golden standards of molecular identification as well as promising alternatives will be discussed. 14.1 Introduction The physiology of the microbial community is fundamental for understanding the processes of anaerobic decomposition of plant material, and has an economic relevance for mankind. The distribution of organisms within the rumen is essential for our understanding of the biochemistry of cellulose degradation (Hungate 1966). A major part of organisms within the rumen fluid encompasses bacteria and flagellates, but fresh and undigested plant material is rapidly colonised by anaerobic fungi. It is now generally known that the degradation of herbal carbohydrates by rumen fungi accelerates the digestion by downsizing the plant tissue particles. Those particles are subsequently more easily decomposed by bacteria and protozoa. The effectiveness of digestion is an important contributor to the health of animals in husbandry (Wulff 2001). M. Eckart, K. Hoffmann, and K. Voigt Institute of Microbiology, School of Biology and Pharmacy, University of Jena, Neugasse 25, 07743 Jena, Germany e-mail: martin.eckart@uni-jena.de K. Fliegerová Department of Biological Basis of Food Quality and Safety, Institute of Animal Physiology and Genetics, Czech Academy of Sciences, v.v.i., Vı́deňská 1083, 14220 Prague 4, Czech Republic Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_14, # Springer-Verlag Berlin Heidelberg 2010 297 298 M. Eckart et al. Because of the economic and scientific interest in this topic, it is not surprising that the first description of “flagellated organisms” living within the rumen was given at the beginning of the twentieth century. But, astonishingly it needed more than 60 years to discover these organisms to be fungi living without any oxygen. The anaerobic environment is mandatory for the ecosystem rumen. It determines the mode of life of microorganisms residing there. Besides being well-known prokaryotes, anaerobic fungi are important producers of short-chain fatty acids, which are an essential source of nutrition for herbivores. Such a unique occupation of a special ecologic niche by a group of heterotrophic, hyphal, and chitin containing eukaryotes inevitably raises the question about the relationships of these fungi. Today, this group is well supported by morphological and molecular data and accepted as the Neocallimastigales (Li et al. 1993). Although the final position within the kingdom Fungi is still unclear, it turned out to be a monophyletic group, as a basal lineage besides or within the phylum Chytridiomycota, and is now recognised as phylum (James et al. 2006). While a flagellated phase through the life cycle of chytridiomycetes is a case sui generis proved for Chytridiales, Rhizophydiales, Spizellomycetales, Blastocladiales, and Neocallimastigales, the rumen fungi are characterised by another unique attribute inside the kingdom Fungi: they live in anaerobiosis. Until now, only several species of gut fungi have been described, probably because of the problematic cultivation and maintenance of these organisms and high morphological variability depending on growth conditions. Extensive studies of a broad range of ruminants and application of modern methods in molecular biology will probably bring deeper insights in microbial communities and species relationships. This survey gives a brief overview of the historical background and of modern trends in species recognition of this interesting fungal group. 14.2 Historical Background and the Discovery of Rumen Fungi Decisive for the terminology of anaerobic gut fungi was a flagellated organism observed within the rumen of herbivores by Liebetanz (1910). This organism was named Callimastix frontalis (Braune 1913) because of its high morphological similarity to Callimastix cyclopis (order Blastocladiales) (Weissenberg 1912), a flagellated parasite of Cyclops. Braune first described the multi-flagellated zoospores seen in Fig. 14.1, but did not recognise them as a stage of a fungal life cycle and misclassified this organism as parasite. The given name C. frontalis led to a number of mis-assignments of parasitic flagellates within this genus. Ultrastructural examinations of C. frontalis by Vavra and Joyón (1966) resulted in the establishment of the new genus Neocallimastix. But unfortunately, the authors did not recognise this organism as a fungus and still considered it as zooflagellate. Eight years later, Whisler et al. (1974) assumed that organisms of the genus Neocallimastix are actually motile spores of an alternate life cycle of Coelomomyces psorophorae – a blastocladialean fungus – and declared the herbivores as 14 Molecular Identification of Anaerobic Rumen Fungi 299 Fig. 14.1 First description of Callimastix as a flagellate parasite. Front view, side view, and cleavage with aequitorial layer (Braune 1913) alternative hosts along with mosquitoes. Orpin (1977) first suggested that these anaerobic organisms living in the rumen actually might be fungi and his assumption was based on the recognition of chitin in the cell walls and on the morphological description of the thallus of different Neocallimastix species (Orpin 1974, 1975, 1976). Orpin’s findings were in contrast to the general belief of microbiologists that no obligate anaerobic fungi can exist and therefore fungal colonies growing in anoxic tubes were discarded as oxygen contaminations (van der Giezen 2002). However, none of these scientists provided a taxonomic definition for Neocallimastix. It was Heath et al. (1983) who linked Neocallimastix to the chytridiomycetes by setting up the new family Neocallimastigaceae within the 300 M. Eckart et al. order Spizellomycetales (phylum Chytridiomycota). The lack of multiple morphological characters has always been and still is a handicap for identifying these organisms within the gut fungi. 14.3 Traditional and Current Systematics Their incapability of locomotion and their appearance resulted in the erroneous classification of fungi as plants before the twentieth century. An own kingdom Fungi was recommended only in 1969 by Whittaker (1969). The Chytridiomycetes, besides the Oomycetes and Hyphochytriomycetes, were the only group of flagellated organisms that shared the class-characteristic cell-wall polymers (Bartnicki-Garcia 1970) and lysine synthetic pathway (Vogel 1964) of the Eumycota, comprising Zygo-, Asco- and Basidiomycetes at that time. In the 1980s, taxonomy and phylogeny of Chytridiomycetes were based on the thallus development, discharge of zoospores, the size, ultrastructural complexity, and organisation of zoospores, as well as number and length of flagella. Furthermore, characteristics like mono- and polycentric development as well as the release of zoospores via diffusion or via papillae affected the taxonomy and phylogeny of these basal fungi (Barr 1978). The anaerobic gut fungi, as a special group of the flagellated fungi, often changed their taxonomic position within the Chytridiomycetes. Because of their late discovery and the unusual physiological character, especially the obligate anaerobiosis, the rumen fungi were placed into different taxonomic groups over time, first into the subdivision Mastigomycotina (Ainsworth 1966), and later into the division Mastigomycota (Alexopoulos and Mims 1979). Within the Mastigomycotina the following zoosporic fungi were accepted: Chytridiomycetes, Hyphochytridiomycetes, Plasmodiophoromycetes, and Oomycetes. The basis for this classification built the zoospore with one or two flagellae as an asexual propagative spore. The class Chytridiomycetes traditionally contained the four orders Chytridiales, Harpochytriales, Blastocladiales, and Monoblepharidales. Some studies mentioned the order Harpochytriales, now known as a synonym of the group Chytridiales (Kirk et al. 2008). Development of molecular genetic methods such as polymerase chain reaction (PCR), cloning, and automated sequencing enabled to generate data for diverse analyses. Traditional phylogeny based on the short-handed phenotypic markers such as morphology, physiology, and biochemistry is now complemented by statistically supported evolutionary analyses, which allowed re-evaluation and re-classification of the whole kingdom Fungi including also the young taxonomic group covering anaerobic fungi. Molecular-biological analysis of gut fungi performed by Li et al. (1993) resulted in the establishment of an own order, the Neocallimastigales, with only one family: Neocallimastigaceae. Recently, Hibbett et al. (2007) postulated a separate phylum for this group, the Neocallimastigomycota, adapted from the paraphyletic origin of the chytridiomycete fungi concluded by James et al. (2006). An informal supertree based on several analyses showed a close relationship to the chytridiomycetes (Hibbett et al. 2007). 14 Molecular Identification of Anaerobic Rumen Fungi 301 Table 14.1 The systematics of the chytridiomycetes based on traditional and modern classification schemes Traditional system Modern system domain Eukaryota domain Eukaryota kingdom Fungi (Linnaeus 1753) Nees 1817 kingdom Fungi (Linnaeus 1753) Nees 1817 phylum Chytridiomycota von Arx 1967 phylum Chytridiomycota von Arx 1967 class Chytridiomycetes (de Bary 1863) class Chytridiomycetes (de Bary 1863) Sparrow Sparrow 1958 1958 order Chytridiales Cohn 1879 order Chytridiales Cohn 1879 order Spizellomycetales Barr 1980 order Spizellomycetales Barr 1980 order Blastocladiales Fitzpatrick 1930 order Rhizophydiales Letcher 2006 order Monoblepharidales Sparrow 1942 order Neocallimastigales Li et al. 1993 class Monoblepharidomycetes Powell 2007 order Monoblepharidales Sparrow 1942 phylum Neocallimastigomycota Powell 2007 class Neocallimastigomycetes Powell 2007 order Neocallimastigales Li et al. 1993 phylum Blastocladiomycota James et al. 2006 class Blastocladiomycetes James et al. 2006 order Blastocladiales Fitzpatrick 1930 However, new phylogenetic approaches display the chytridiomycetes again as monophyletic group (Ebersberger et al. 2010). Therefore, the separate phylum Neocallimastigomycota seems to be redundant. A comparison of both classical taxonomy, based on morphology and physiology, and modern systematic methods, based on up-to-date molecular-genetic techniques, is shown in Table 14.1. At present, the family Neocallimastigaceae comprises six genera1 (Adl et al. 2005): Anaeromyces (Breton et al. 1990), Caecomyces (Gold et al. 1988), Cyllamyces (Ozkose et al. 2001), Neocallimastix (Vavra and Joyon ex Heath 1983), Orpinomyces (Barr et al. 1989), and Piromyces (Gold et al. 1988). An overview of the six taxonomic groups within the family Neocallimastigaceae is shown in Table 14.2. 14.4 Phylogeny Traditional phylogenetic results supported by molecular-genetic data can redraw evolutionary hypothesis and consequently the affinity of organisms to taxonomic groups. Like phenotypic characterisations, molecular phylogenetics should never be based just on one character. Comparisons or combinations of morphological and genetic characters lead to stable and well supported evolutionary hypotheses and with this strengths and weaknesses of genetic markers become obvious. A marker of high diagnostic value has to be unique to a species or even to a strain and at the 1 http://indexfungorum.org/Names/familyrecord.asp?strRecordID=81063 302 M. Eckart et al. Table 14.2 Survey of the species from the anaerobic chytrids Genus Species Author Neocallimastix frontalis (RA Braune) Vavra and Joyón 1966 ex IB Heath et al. (1983) hurleyensis Theodorou and Webb (1991) joyonii Breton, Gaillard, Bernalier, Bonnemoy and Fonty (1988) patriciarum Orpin and Munn (1986) variabilis Ho and Barr (1993) Anaeromyces elegans Ho (1993) mucronatus Breton et al. (1990) Caecomyces communis Gold et al. (1988) equi Gold (1988) sympodialis Chen, Tsai and Chien (2007) Cyllamyces aberensis Ozkose et al. (2001) Orpinomyces bovis Barr et al. (1989) intercalaris Ho et al. (1994) (Breton, Bernalier, Bonnemoy, Fonty, Gaillard and Gouet) joyonii1 Li, Heath and Cheng (1990)) Piromyces citronii Gaillard, Breton, Dusser and Julliand (1995) communis Gold, Heath, and Bauchop (1988) dumbonicus Li (1990) mae Li (1990) minutus Ho (1993) polycephalus Chen, Chien and Hseu (2002) rhizinflatus Breton, Dusser, Gaillard, Guillot, Millet and Prensier (1991) spiralis Ho (1993) 1 Basionym, current name: Neocallimastix joyonii Breton, Bernalier, Bonnemoy, Fonty, B. Gaillard & Gouet 1989 same time ubiquitous for all taxa. The more various the set taxonomic groups is, the more conserved the marker has to be. On the other hand, clustering on lower level, starting with the family, requires more variable data to distinguish between species or even strains (outlined in Fig. 14.2). Clustering methods use differences between partitions of given data to rebuild cladistic relationships. To get quality estimation such as bootstrap proportions (BTs or BP), checking the robustness of a set of data (Felsenstein 1985) is required. Highly conserved data lead to stable reconstructions in early branches caused by low clade stability supports. An example is given in Fig. 14.2. There is no possibility to distinguish between taxon2 and taxon3 based on identical sequence data. This exemplary marker is not adequate for molecular diagnostics on a lower taxonomic level such as genus or species. We demonstrate these problems with an analysis based on real data in Fig. 14.3. The gene encoding actin is highly conserved in eukaryotes. The coding sequence divergence between plant and non-plant actin genes shows only 15% or less variability (Hightower and Meagher 1986). Therefore, this marker demonstrates perfectly the relationship between anaerobic and aerobic chytridiomycetes, with the zygomycetous order Mucorales as outgroup (Fig. 14.3). Varieties of species level cannot be determined with actin data, as this marker lacks molecular diagnostics possibilities. 14 Molecular Identification of Anaerobic Rumen Fungi 303 Fig. 14.2 Schematic illustration of problems occurring during the application of single phylogenetic markers. The master sequence should be ATTGCTAAGCGA; the (recent) taxa show modified sequences. Changes compared to the consensus sequences are colour-coded in red. Occurring problems are obvious: a stable backbone with statistical support is only possible with data that are not highly diverse. However, differentiation at higher branches requires variable sequences. To combine these datasets, several approaches like supermatrix or supertree methods can be applied Nevertheless, highly variable data could result in an unstable reconstruction of early branches caused by “long-branch-attraction” (Bergsten 2005). Although high variable data could help measure distances between the closest neighbours and other taxa on lower taxonomic levels, the variability of the data could lead to false positive congruence, such as analogy instead of homology. An example is shown in Fig. 14.4. The high variability of the internal transcribed spacer (ITS) sequences of chytridiomycetes allow to distinguish even between strains, but alternative ways to cluster these data decrease the robustness of the data set. One problem is the differentiation of homologous and paralogous markers. Homology is not a problem if orthologous genes are involved, but paralogous genes can lead to misinterpreted results, similar to the comparison of “apples and oranges”. One example would be the eukaryotic translation elongation factor 1-a (EF-1a) with more than one copy within the genomes of fungi (see fungal genomes published by the JGI at http:// genome.jgi-psf.org/). False positive results in phylogenetic analysis based on alignments of paralogous genes are not always obvious as shown in Fig. 14.5. 14.5 Predicted Impact of Molecular Markers on Future Identification and Phylogeny Based on the theory of evolution, highly conserved genes of diverse taxonomic groups could be amplified by the combination of PCR techniques and universal oligonucleotides (primers). The most commonly used DNA region for moleculargenetic phylogeny is the highly repetitive cluster of the nuclear ribosomal DNA (rDNA). The nucleotide sequences of the nuclear small (SSU) and large (LSU) 304 M. Eckart et al. Fig. 14.3 Phylogeny of the Neocallimastigomycota and other chytridiomycetes based on a Maximum Likelihood analysis of actin sequences from 24 taxa with a total of 903 aligned characters (unpublished sequences). (methodical informations: GAMMA þ P-Invar model with RAxML 7.0.4 GTR-CAT (rapid hill-climbing bootstrapping method (Stamatakis 2006, 2008)), 10,000 rapid bootstrap inferences before a thorough ML search; final ML Optimization Likelihood: – 4676.139464) 14 Molecular Identification of Anaerobic Rumen Fungi Fig. 14.4 Phylogeny of the Neocallimastigomycota and other chytridiomycetes based on a Maximum Likelihood analysis of ITS sequences from 49 taxa with a total of 1,161 aligned characters (unpublished sequences). (methodical informations: GAMMA þ P-Invar model with RAxML 7.0.4 GTR-CAT, (Stamatakis 2006, 2008), 10,000 rapid bootstrap inferences before a thorough ML search. Final ML Optimization Likelihood: – 17766.020153) 305 306 M. Eckart et al. Fig. 14.5 Phylogeny of the Neocallimastigomycota and other chytridiomycetes based on a Maximum Likelihood analysis of tef sequences from 21 taxa with a total of 1,377 aligned characters (unpublished sequences). Included paralogous copies disturb the correct species assignments. (methodical informations: GAMMA þ P-Invar model with RAxML 7.0.4 GTR-CAT (Stamatakis 2006, 2008), 10,000 rapid bootstrap inferences before a thorough ML search. Final ML Optimization Likelihood: – 5723.672908) 14 Molecular Identification of Anaerobic Rumen Fungi 307 subunits are separated by the non-coding DNA sequences of the internal transcribed spacer (ITS) 1 and 2 and the non-transcribed intergenic spacer (IGS). Lacking a sufficient evolutionary pressure, the non-coding regions allow the separation of organisms down to the levels of species and strains. Using the flanking conserved sequences of 18S (SSU) and 28S (LSU) rDNA, these regions can easily be amplified with universal primers. Unfortunately, the ITS regions are not single copy regions. Although the ribosomal DNA cluster follows concerted evolution (Arnheim 1983), the intra-specific variability among organisms cannot be denied. Usage of this region as molecular barcode marker is therefore questionable, especially if there is no reliable and supporting approach for species identification based on e.g. morphology (Nilsson et al. 2008). Moreover, the variability of the ITS region is sometimes not high enough to separate at the level of species as shown for the fungal genus Penicillium (Skouboe et al. 1996, 1999). This experience enforced the search and establishment of alternative genetic markers like the introncontaining protein coding genes actin (act), eukaryotic translation elongation factor 1-a (tef), or beta-tubulin (btub). A profound base for this approach requires reference strains, which need to be morphologically and genetically well characterised, and also the subsequent completion of the published results and sequence submissions. First efforts to identify the anaerobic gut fungi by molecular genetic methods were done by Doré and Stahl (1991) and Bowman et al. (1992). Their approaches relied on partial 18S rDNA sequences for including the anaerobic fungi into the chytridiomycetes, but the authors did not separate the species within the genera (Doré and Stahl 1991; Bowman et al. 1992). Trying to clarify the phylogenetic relationships within the order Neocallimastigales using sequence analysis (ITS1) combined with morphological features, ultrastructures and mitotic characters have led to seperation of the order Neocallimastigales (Li et al. 1993). Isozyme analyses or DNA hybridisation has also been used with the aim to clarify identification of anaerobic gut fungi and to increase the level of specificity (Ho et al. 1994). A fast and easy method for the differentiation of polycentric anaerobic fungi is available by (RFLP) analysis of ITS spacer and/or fragments of ribosomal large subunit (28S rDNA) digested by proper endonucleases. However, the ribosomal small subunit (18S rDNA) turned out to be too conservative to get a well resolved DNA polymorphism, and therefor is not very suitable for this type of analysis (Fliegerová et al. 2006). Methods of molecular biology are very promising, but “old-fashioned” taxonomy is still substantiated despite many discrepancies. The classical approach of Neocallimastigales identification is based on their morphological characters. Thallus shape (filamentous or bulbous), zoosporangial development (monocentric or polycentric), and number of flagella per zoospore (uni- or polyflagellated) are decisive for genus differentiations, while the ultrastructure of the zoospore is determinative for species. (Heath et al. 1983; Orpin and Munn 1986; Munn et al. 1988; Webb and Theodorou 1991). Unfortunately, characters observable by light microscopy vary with culture conditions and are highly pleomorphic (Brookman et al. 2000). Moreover, the cultures often fail to produce important structures 308 M. Eckart et al. (sporangia and zoospores) making identification even more problematic. Also the differentiation of species using ultrastructural features of the zoospores is questionable, because ultrastructure depends not only on the age of microorganisms but also on the method and quality of their preparation (Ho and Barr 1995). 14.6 Molecular Identification and DNA Barcoding One of the important characteristics of anaerobic fungi is their flagellated stage in life cycle. However, flagellated zoospores, can be found also in other aquatic fungi, like the Blastocladiomycota and the Chytridiomycota sensu stricto and also in some protists, e.g., the stramenopiles and among those the oomycetes, which are derived brown algae. The flagellae of these organisms caused the mis-applications of taxonomic and phylogenetic assignments as it happened to Braune with Neocallimastix (1913). The elucidation of morphological characters is valuable and indispensable, but has to be supported by techniques of molecular genetics because the pleomorphic shape of fungi leads to complications in their identification. Therefore, molecular information becomes more and more important as a primary source for species recognition. Now, 90 years after the discovery of the anaerobic rumen fungi, molecular phylogenetic studies confirmed their relationship to the kingdom Fungi (Förster et al. 1990; Bowman et al. 1992). The choice of molecular genetic markers in the kingdom Fungi, respectively the phylum Chytridiomycota, is clearly arranged. Today, state of the art comprises seven markers for fungal phylogeny that provides data over the complete spectrum of the kingdom: 18S rDNA, 28S rDNA, ITS1 and ITS2 including the 5.8S rDNA, rpb1, rpb2, tef, and beta-tubulin. In special cases like pathogenic species or organisms of industrial importance, some additional markers exhibiting a higher specificity were developed. Such markers encompass not only genes encoding calmodulin, Mcm7 (MS456), and Tsr1 (MS277) (Aguileta et al. 2008; Schmitt et al. 2009) but also physiological properties such as toxins or extrolite profiles, which are well established, for example, the ascomycetous genus Penicillium (Samson and Frisvad 2004). To find the most useful marker for “tagging” all forms of life is the aim of many current projects involved in “DNA barcoding”. DNA barcoding is an approach to identify any organism based on sequence analysis of selected genomic regions. Access to these regions should be as universal as possible, comparable, reproducible, and relatively easy to accomplish. Barcoding is thought to serve not only the identification or verification of known specimens but also to contribute in the discovery of new, undescribed species. Although DNA barcoding already proved to be a very useful tool for the discovery of cryptic species, which are by definition not differentiable by morphological features (Hebert et al. 2004), barcoding is nevertheless error-prone. Depending on the method used, DNA barcoding turned out to be not always sufficient for species recognition (Brower 2006; DeSalle et al. 2005; Whitworth et al. 2007). One of the major problems in all barcoding 14 Molecular Identification of Anaerobic Rumen Fungi 309 a b Phylum Ascomycota Basidiomycota environmental samples Glomeromycota Microsporidia unclassified fungi Fungi incertae sedis Zoopagomycotina Entomophthoromycotina Kickxellomycotina Mucoromycotina unkown Chytridiomycota Neocallimastigomycota Blastocladiomycota CoS 1,124,132 197,974 41,955 11,698 5,676 5,107 5,092 32 439 276 3,948 397 1,683 465 134 Fig. 14.6 Schematic illustration and number of nucleotide sequences provided by the International Nucleotide Sequence Database Collaboration. (e.g., Genbank). (a) Percentages of the number of fungal sequences provided in GenBank. The graph shows the total number of submitted sequences within the kingdom Fungi. The subparts describe the single phyla based on the taxonomy provided by the TaxBrowser at NCBI. The group Dikarya is represented by approximately 96% of all available sequences. The number of sequences that were generated of environmental samples is higher than that of all other phyla with the exception of the Dikarya. The Neocallimastigomycota represent with 465 sequences the second smallest group of fungal organisms represented as nucleotide sequences within GenBank. (b) Survey of the nucleotide sequences provided by the International Nucleotide Sequence Database Collaboration (as of May 1st, 2009) approaches is still the question which molecular tool should be used, since every further step in species identification is based on it. In animal systems, the mitochondrial cox1 is widely applied (Hebert et al. 2003), although its sufficiency is already questioned (Goetze 2003). With a slower evolutionary rate of this cytochrome c oxidase, this marker is not applicable for flowering plants (Kress et al. 2005). One of the most discussed marker in fungal taxonomy and phylogenetics is still the ITS region with all its afore mentioned advantages and disadvantages (see Sect. 5). 310 M. Eckart et al. Nevertheless, provided sequence data usable for species identification for anaerobic gut fungi are restricted, e.g., only 163 “ITS” tagged sequences are assigned to the order Neocallimastigales (465 nucleotide sequences in summary, compare Fig. 14.6), with only 17 sequences assigned to full taxon names. Because of missing mitochondria in anaerobic fungi (hydrogenosomes instead), mitochondrial based barcode markers are out of question (Bullerwell and Lang 2005). The need for a complete barcoding database, as always demanded (Ekrem 2007), is obvious. But another major drawback is the data deposited in such a barcode database. An adequate number of well-defined reference specimens are a prerequisite for species identification and especially for species discovery. Such references should encompass all possible variances within defined species boundaries, e.g., geographically based variations (DeSalle et al. 2005; Meyer and Paulay 2005). Originally thought to be a fast, cheap, and easy-to-access method for the assignment of “unknown” to “known” specimens, molecular barcoding should be used with caution. On the one hand, supplementing a barcode marker with additional information about e.g., morphology, biogeography, or even more molecular data will miss the aim of a single easy-to-use marker for species assignment. But on the other hand, supplementing data is necessary as a specimen cannot be identified or described with certainty by one molecular attribute (Brower 2006; Will et al. 2005). Storing new data in a database is always tied with responsibility of the submitter. Open-access to such databases is necessary but at the same time prone to errors and losing its value as shown by GenBank at the NCBI (Bridge et al. 2003). 14.7 Conclusion and Future Line of Research According to the efforts of Aguileta et al. (2008) and Schmitt et al. (2009) more alterantive barcoding markers need to be established and validated in order to get a reliable identification which is in concordance with morphological and ultrastructural characters. The increase of the complexity of research on anaerobic rumen fungi in their composite ecosystems requires a common platform for strain and data shared among the scientific community. It is necessary to gain a certain homogeneity and common use of reference and type strains including reference sequences of barcode markers and other characters suitable for a reliable identification of anaerobic rumen fungi. This is a fundamental for cultivation-independent detection in the natural ecosystems and habitats of anaerobic fungi as performed by Fliegerová et al. (2010). Acknowledgments The authors express their gratitude to L. Jay Yanke (Agriculture and AgriFood Canada, Lethbridge Research Centre, Lethbridge, AB, Canada) for providing Neocallimastix frontalis strain SR4. This project was a component of the institutional research plan (AV OZ 5045 0515) of the Institute of Animal Physiology and Genetics, Academy of Sciences of the Czech Republic in Prague. The czech team was supported by The National Agency for Agriculture Research (project no. QI92A286/2008), The german team was supported by the Deutsche Forschungsgemeinschaft (project no. Vo 772/7-1), which is part of a bilateral grant between the Czech Science Foundation and the Deutsche Forschungsgemeinschaft. 14 Molecular Identification of Anaerobic Rumen Fungi 311 References Adl SM, Simpson AGB, Farmer MA, Andersen RA, Anderson OR, Barta JR, Bowser SS, Brugerolle G, Fensome RA, Fredericq S, James TY, Karpov S, Kugrens P, Krug J, Lane CE, Lewis LA, Lodge J, Lynn DH, Mann DG, McCourt RM, Mendoza L, Moestrup O, MozleyStandridge SE, Nerad TA, Shearer CA, Smirnov AV, Spiegel FW, Taylor MFJR (2005) The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. J Eukaryot Microbiol 52:399–451 Aguileta G, Marthey S, Chiapello H, Lebrun MH, Rodolphe F, Fournier E, Gendrault-Jacquemard A, Giraud T (2008) Assessing the performance of single-copy genes for recovering robust phylogenies. Syst Biol 57:613–627 Ainsworth GC (1966) A general purpose classification of the fungi. Bibl System Mycol 4:1–4 Alexopoulos CJ, Mims CW (1979) Introductory Mycology 38:1–613 Arnheim N (1983) Concerted evolution of multigene families. In: Nei M, Koehn RK (eds) Evolution of genes and proteins. Sinauer Associated, Sunderland, MA, pp 38–61 Barr DJS (1978) Taxonomy and phylogeny of chytrids. Biosystems 10:153–165 Barr DJS, Kudo H, Jakober KD, Cheng KJ (1989) Morphology and development of rumen fungi: Neocallimastix sp., Piromyces communis, and Orpinomyces bovis gen.nov., sp.nov. Can J Bot 67:2815–2824 Bartnicki-Garcia S (1970) Cell wall composition and other biochemical markers in fungal phylogeny. In: Harborne JB (ed) Phytochemical phylogeny. Academic Press, New York, pp 81–103 Bergsten J (2005) A review of long-branch attraction. Cladistics 21(2):163–193 Bowman BH, Taylor JW, Brownlee AG, Lee J, Lu S, White TJ (1992) Molecular evolution of the fungi: relationship of the basidiomycetes, ascomycetes, and chytridiomycetes. Mol Biol Evol 9:285–296 Braune R (1913) Untersuchungen über die im Wiederkäuermagen vorkommenden Protozoen. Arch Protistenk 32:111–170 Breton A, Bernalier A, Dusser M, Fonty G, Gaillard-Martinie B, Guillot J (1990) Anaeromyces mucronatus nov. gen., nov. sp. A new strictly anaerobic rumen fungus with polycentric thallus. FEMS Microbiol Lett 70:177–182 Bridge PD, Roberts PJ, Spooner BM, Panchal G (2003) On the unreliability of published DNA sequences. New Phytol 160:43–48 Brookman JL, Mennim G, Trinci AP, Theodorou MK, Tuckwell DS (2000) Identification and characterization of anaerobic gut fungi using molecular methodologies based on ribosomal ITS1 and 18S rRNA. Microbiology 146:393–403 Brower AVZ (2006) Problems with DNA barcodes for species delimitation: ‘ten species’ of Astraptes fulgerator reassessed (Lepidoptera: Hesperiidae). Syst Biodivers 4:127–132 Bullerwell CE, Lang BF (2005) Fungal evolution: the case of the vanishing mitochondrion. Curr Opin Microbiol 8:362–369 DeSalle R, Egan MG, Siddall M (2005) The unholy trinity: taxonomy, species delimtiation and DNA barcoding. Philos Trans R Soc Lond B Biol Sci 360:1905–1916 Doré J, Stahl DA (1991) Phylogeny of anaerobic rumen Chytridiomycetes inferred from small subunit ribosomal RNA sequence comparisons. Can J Bot 69:1964–1971 Ebersberger I, Strauss S, Kupczok A, Kothe E, von Haeseler A, Gube M, Eckart M, Voigt K (2010) A stable backbone for the fungi. in revision Ekrem T, Willassen E, Stur E (2007) A comprehensive DNA sequence library is essential for identification with DNA barcodes. Mol Phylogenet Evol 43:530–542 Felsenstein J (1985) Confidence limits of phylogenies: an approach using the bootstrap. Evolution 39:783–791 Fliegerová K, Mrázek J, Voigt K (2006) Differentiation of anaerobic polycentric fungi by rDNA PCR-RFLP. Folia Microbiol (Praha) 51:273–277 312 M. Eckart et al. Fliegerová K, Mrázek J, Hoffmann K, Zábranská J, Voigt K (2010) Diversity of anaerobic fungi within cow manure determined by ITS1 analysis. Folia Microbiol (Praha): in press Förster H, Coffey MD, Elwood H, Sogin ML (1990) Sequence analysis of the small subunit ribosomal RNAs of three zoosporic fungi and implications for fungal evolution. Mycologia 82:306–312 Goetze E (2003) Cryptic speciation on the high seas; global phylogenetics of the copepod family Eucalanidae. Proc R Soc B 270:2321–2331 Gold JJ, Heath IB, Bauchop T (1988) Ultrastructural description of a new chytrid genus of caecum anaerobe, Caecomyces equi gen. nov., sp. nov., assigned to the Neocallimasticaceae. BioSystems 21:403–415 Heath IB, Bauchop T, Skipp RA (1983) Assignment of the rumen anaerobe Neocallimastix frontalis to the Spizellomycetales (Chytridiomycetes) on the basis of its polyflagellate zoospore ultrastructure. Can J Bot 61:295–307 Hebert PDN, Ratnasingham S, deWaard JR (2003) Barcoding animal life: cytochrome c oxidase subunit 1 divergences among closely related species. Proc R Soc B 270(Suppl):96–99 Hebert PD, Penton EH, Burns JM, Janzen DH, Hallwachs W (2004) Ten species in one: DNA barcoding reveals cryptic species in the neotropical skipper butterfly Astraptes fulgerator. Proc Natl Acad Sci USA 101:14812–14817 Hibbett DS, Binder M, Bischoff JF, Blackwell M, Cannon PF, Eriksson OE, Huhndorf S, James T, Kirk PM, Lücking R, Lumbs HT, Lutzoni F, Matheny PB, McLaughlin DJ, Powell MJ, Redhead S, Schoch CL, Spatafora JW, Stalpers JA, Vilgalys R, Aime MC, Aptroot A, Bauer R, Begerow D, Benny GL, Castlebury LA, Crous PW, Dai YC, Gams W, Geiser DM, Griffith GW, Gueidan C, Hawksworth DL, Hestmark G, Hosaka K, Humber RA, Hyde KD, Ironside JE, Kõljalg U, Kurtzman CP, Larsson KH, Lichtwardt R, Longcore J, Miadlikowska J, Miller A, Moncalvo JM, Mozley-Standridge S, Oberwinkler F, Parmasto E, Reeb V, Rogers JD, Roux C, Ryvarden L, Sampaio JP, Schüssler A, Sugiyama J, Thorn RG, Tibell L, Untereiner WA, Walker C, Wang Z, Weir A, Weiss M, White MM, Winka K, Yao YJ, Zhang N (2007) A higher-level phylogenetic classification of the fungi. Mycol Res 111:509–547 Hightower RC, Meagher RB (1986) The molecular evolution of actin. Genetics 114:315–332 Ho Y, Barr D (1995) Classification of anaerobic gut fungi from herbivores with emphasis on rumen fungi from malaysia. Mycologia 87:655–677 Ho YW, Khoo IY, Tan SG, Abdullah N, Jalaludin S, Kudo H (1994) Isozyme analysis of anaerobic rumen fungi and their relationship to aerobic chytrids. Microbiology 140:1495–1504 Hungate RE (1966) The rumen and its microbes. Academic Press, New York James TY, Letcher PM, Longcore JE, Mozley-Standridge SE, Porter D, Powell MJ, Griffith GW, Vilgalys R (2006) A molecular phylogeny of the flagellated fungi (Chytridiomycota) and a proposal for a new phylum (Blastocladiomycota). Mycologia 98:860–871 Kirk PM, Cannon PF, Minter DW, Stalpers JA (2008) Dictionary of the fungi, 10th edn. CABI Publishing, Wallingford, UK Kress WJ, Wurdack KJ, Zimmer EA, Weigt LA, Janzen DH (2005) Use of DNA barcodes to identify flowering plants. Proc Natl Acad Sci USA 102:8369–8374 Li J, Heath IB, Packer L (1993) The phylogenetic relationships of the anaerobic chytridiomycetous gut fungi (Neocallimasticaceae) and the Chytridiomycota II. Cladistic analysis of structural data and description of Neocallimasticales ord. nov. Can J Bot 71:393–407 Liebetanz E (1910) Die parasitischen Protozoen des Wiederkäuermagens. Arch Protistenk 19:19–90 Meyer CP, Paulay G (2005) DNA barcoding: error rates based on comprehensive sampling. PLoS Biol 3:e422 Munn EA, Orpin CG, Greenwood CA (1988) The ultrastructure and possible relationships of four obligate anaerobic chytridiomycete fungi from the rumen of sheep. Biosystems 22:67–81 Nilsson RH, Kristiansson E, Ryberg M, Hallenberg N, Larsson K (2008) Intraspecific ITS variability in the kingdom fungi as expressed in the international sequence databases and its implications for molecular species identification. Evol Bioinform 4:193–201 14 Molecular Identification of Anaerobic Rumen Fungi 313 Orpin CG (1974) The rumen flagellate Callimastix frontalis: does sequestration occur? J Gen Microbiol 84:395–398 Orpin CG (1975) Studies on the rumen flagellate Neocallimastix frontalis. J Gen Microbiol 91:249–262 Orpin CG (1976) Studies on the rumen flagellate Sphaeromonas communis. J Gen Microbiol. 94:270–280 Orpin CG (1977) Invasion of plant tissue in the rumen by the flagellate Neocallimastix frontalis. J Gen Microbiol 98:423–430 Orpin C, Munn E (1986) Neocallimastix patriciarum: a new member of the Neocallimasticaceae inhabiting the sheep rumen. Trans Br Mycol Soc 86:178–181 Ozkose E, Thomas BJ, Davies DR, Griffith GW, Theodorou MK (2001) Cyllamyces aberensis gen. nov sp.nov., a new anaerobic gut fungus with branched sporangiophores isolated from cattle. Can J Bot 79:666–673 Samson RA, Frisvad JC (eds) (2004) Penicillium subgenus Penicillium: new taxonomic schemes, mycotoxins and other extrolites. Stud Mycol 49:1–251 Schmitt I, Crespo A, Divakar PK, Fankhauser JD, Herman-Sackett E, Kalb K, Nelsen MP, Nelson NA, Rivas-Plata E, Shimp AD, Widhelm T, Lumbsch HT (2009) New primers for promising single copy genes in fungal phylogenetics and systematics. Persoonia 23:35–40 Skouboe P, Boysen M, Pedersen LH, Frisvad JC, Rossen L (1996) Identification of Penicillium species using the internal transcribed spacer (ITS) regions. In: Rossen L, Rubio V, Dawson MT, Frisvad JC (eds) Fungal identification techniques. European Commission, Brussels, Belgium, pp 160–164 Skouboe P, Frisvad JC, Taylor JW, Lauritsen D, Boysen M, Rossen L (1999) Phylogenetic analysis of nucleotide sequences from the ITS region of terverticillate Penicillium species. Mycol Res 103:873–881 Stamatakis A (2006) RAxML-VI-HPC: Maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22(21):2688–2690 Stamatakis A, Hoover P, Rougemont J (2008) A rapid bootstrap algorithm for the RAxML webservers. Systematic Biology 75(5):758–771 van der Giezen M (2002) Strange fungi with even stranger insides. Mycologist 16:129–131 Vavra J, Joyón L (1966) Etude sur la morphologie, le cycle évolutif et la position systématique de Callimastix cyclopis Weissenberg, 1912. Protistologica 2:5–16 Vogel HJ (1964) Distribution of lysine pathways among fungi: evolutionary implications. Am Nat 98:435 Webb J, Theodorou MK (1991) Neocallimastix hurleyensis sp.nov., an anaerobic fungus from the ovine rumen. Can J Bot 69:1220–1224 Weissenberg S (1912) Sitzungsbericht der Gesellschaft naturfoschender Freunde zu Berlin 5:299 Whisler HC, Zebold SL, Shemanchuk JA (1974) Alternate host for mosquito parasite Coelomomyces. Nature 251:715–716 Whittaker RH (1969) New concepts of kingdoms or organisms. Evolutionary relations are better represented by new classifications than by the traditional two kingdoms. Science 163:150–160 Whitworth TL, Dawson RD, Magalon H, Baudry E (2007) DNA barcoding cannot reliably identify species of the blowfly genus Protocalliphora (Diptera: Calliphoridae) Proc Bio Sci 274:1731–1739 Will KW, Mishler BD, Wheeler QD (2005) The perils of DNA barcoding and the need for integrative taxonomy. Syst Biol 54:844–851 Wulff C (2001) Untersuchungen zum Thiamin- und Thiaminderivatgehalt im Pansensaft des Rindes nach Verfütterung von mit Mucor racemosus Fresenius verpilztem Heu (in vitro), Thesis, Tierärztliche Hochschule Hannover, Hannover Part II Human Pathological and Clinical Aspects Chapter 15 New Approaches in Fungal DNA Preparation from Whole Blood and Subsequent Pathogen Detection Via Multiplex PCR Roland P.H. Schmitz, Raimund Eck, and Marc Lehmann Abstract Sepsis is a life-threatening disease that results from excessive host responses to microbial infections. Fungal pathogens mainly contribute to lethal outcomes and high treatment budgets. Numerous trials revealed that the mortality rates of septic patients could be reduced if appropriate anti-infective approaches are promptly initiated. This demands a forthwith identification of the causative pathogen(s) and antibiotic resistances. However, standard procedures (e.g., blood cultures) deliver first results after 2–3 days. Facing the time-to-result for cultural pathogen detection, culture independent nucleic acid amplification techniques (NAT) are increasingly desirable to deliver a reliable basis for a targeted antibiotic regimen within the first decisive hours of the disease. Crucial steps in the detection of pathogens within whole blood concern cell lysis and the disproportion of pathogen and human background DNA Standard analytical methods applied and current developments in sepsis diagnostics are reviewed. New tools are introduced which accelerate the clinical investigation course and improve the sensitivity as well as the quality of NAT-based genus and species detections. 15.1 Introduction: Fungi as Sepsis Causative Pathogens Life-threatening fungal and bacterial infections and their outcomes – sepsis and consecutive organ failure – are frequent complications of hospitalized patients, with an increasing number of 18 million new sepsis cases each year worldwide and with a mortality rate of 30–50% (Slade et al. 2003). Sepsis results from the hosts response to fungal and bacterial (and protozoan) infections, whereas a malfunction of the defence and repair system is responsible for the development of organ dysfunctions and at last multiorgan failure. R.P.H. Schmitz, R. Eck, and M. Lehmann SIRS-Lab GmbH, Winzerlaer Str. 2, 07745 Jena, Germany e-mail: schmitz@sirs-lab.com Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_15, # Springer-Verlag Berlin Heidelberg 2010 317 318 R.P.H. Schmitz et al. In Germany, about 60,000 out of 1,54,000 patients die from severe sepsis, which therefore is one of the most frequent causes of death in intensive care units (ICU). The incidence of sepsis has risen due to the use of invasive devices, aging of the population, and the higher incidence of immunosuppressive conditions such as chemotherapy for cancer and acquired immunodeficiency syndrome (AIDS) (Martin et al. 2003). About 30% of the intensive medicine budget is expended for the treatment of those patients (Reinhart et al. 2006). On average, the treatment of candidaemia amounts to about US $44,536 for each patient, mainly attributed to prolonged hospitalization (Pfaller et al. 2005), whereas treatment of sepsis costs only US $22,000 per patient in the USA (US $16.7 billion spent each year for sepsis care; Angus and Wax 2001), which marks the discrepancies in care expenses of septical infections caused by fungal and bacterial pathogens. From the 1970s to 1990s, septical infections in Germany were attributed mainly to Gram-negative bacterial species, whereas currently Gram-positives, fungi, and multi-infections are rising (Bauer et al. 2006; Karlowsky et al. 2004; Martin et al. 2003). The incidence of fungal species in septical infections was determined to be mainly contributed by Candida spp. (Fig. 15.1), especially C. albicans, with prominent findings in several clinical studies: compared to bacterial sepsis causative pathogens, the incidence of fungal species was quite significant with Candida spp. Fig. 15.1 Frequencies of fungal sepsis causative pathogens as determined in 11 clinical studies. Individual species were designated only in a part of the studies, which accounts for their percentage distribution as individual species or fungi/Candida sp. The boxplot data were calculated using R (R Development Core Team, 2007). The data were taken from: Bodmann and Vogel (2001); Cockerill et al. (1997); Focht and Adam (2004); Fridkin and Gaynes (1999); Geerdes et al. (1992); Geffers et al. (2004); Jones et al. (2004); Karlowsky et al. (2004); Kübler et al. (2004); Styers et al. (2006); Vincent et al. (2006); R: A language and environment for statistical computing. R Foundation for Statistical Computing (2007) Vienna, Austria, ISBN 3-900051-07-0 15 New Approaches in Fungal DNA Preparation from Whole Blood 319 10%, Aspergillus spp. 5–15%, and Fusarium spp. < 2% (Richardson and Kokki 1999; Vincent et al. 1998; Duthie and Denning 1995; Gamis et al. 1991). Candida albicans was then identified to be the fifth and seventh frequent causative pathogen of nosocomial infections in Germany and the USA, respectively (Geffers et al. 2004; Fridkin and Gaynes 1999), while candidaemia in general accounts for the fourth frequent nosocomial infection affecting the blood circuit (Liu et al. 2006; Gudlaugsson et al. 2003; Pappas et al. 2003). The spectrum of nosocomial fungal pathogens in particular comprises mainly yeasts – first of all C. albicans (>50%), followed by non-C. albicans Candida spp. (~45%, e.g., C. glabrata, C. krusei, C. parapsilosis, C. tropicalis) as well as molds (2–5%, e.g., Aspergillus spp.). The lethality of nosocomial C. albicans infections was quoted to be 22–40% (Gudlaugsson et al. 2003; Nolla-Salas et al. 1997; Fraser et al. 1992; Wey et al. 1988). The incidence of sepsis and the number of sepsis-related deaths were increasing in the USA, although the overall mortality rate among patients with sepsis was declining: the rate of sepsis due to fungal organisms increased by 207% between 1979 and 2000 (with Gram-positive bacteria becoming the predominant pathogens after 1987) (Martin and Mannino 2003). Candidaemia occurrence was proven to be associated with an extraordinary high mortality rate among critically ill patients: in fungal infections the patients mortality was >60% (85% and 45.2% in medical and surgical patients, respectively) (Charles et al. 2003). Causing identical clinical syndromes in severe sepsis/septic shock, the mortality of patients with bacterial infections was only determined to be 30–50%. About 70% of all fungal infections were diagnosed post mortem (Schmidt et al. 1991). 15.2 Sepsis Diagnostics: An Overview of Current Methods Given the high mortality rates from patients suffering from candidaemia, there is an obvious need for more accurate and rapid identification strategies. Early detection of systemic fungal infection and prompt and adequate antibiosis in the first few hours of infection are the crucial steps for an effective sepsis therapy (GarnachoMontero et al. 2003; Fine et al. 2002; Ibrahim et al. 2000). Epidemiological data confirm that a doubling of mortality is the consequence of inadequate therapies (Vallés et al. 2003) (Fig. 15.2) and an increase of mortality of more than ~7% per hour is proven in cases of delayed adequate (directed) antibiotic treatment in case of sepsis, irrespective of the causative pathogen (Iregui et al. 2002). Unsurprisingly, an increase of mortality has also been proven for delayed empirical (nondirected) therapies in cases of systemic candidiasis (Morrell et al. 2005). Diagnosis of invasive fungal infections (IFI) is generally based on direct pathogen detection in body fluids and in aseptically withdrawn samples of the skin, lung, liver, and CNS biopsy, respectively, via growth cultures. A histological proof within inflamed tissue is also accepted (Ascioglu et al. 2002). However, the conventional identification of pathogenic fungi in clinical microbiology 320 R.P.H. Schmitz et al. Fig. 15.2 Reduction of mortality by adequate antibiotic therapy laboratories bases on phenotypic features and physiological tests and remains a significant problem in general. The clinical symptoms are difficult to interpret and the findings of noninvasive methods (e.g., computed tomographic scanning and X-ray) are not specific (Richardson and Kokki 1999). Deep-tissue sample cultures from infections with focal lesions are frequently negative (Vincent et al. 1998; Duthie and Denning 1995). Direct microscopy and histopathological examination are rapid, but they do not always allow identification of the infecting agent to the species level (Richardson and Kokki 1999; Vincent et al. 1998). Even though monoclonal antibody-based enzyme-linked immunosorbent assays (ELISAs) for circulating Aspergillus and Candida antigens (e.g., Cand-Tec1 latex agglutination assay, Ramco Laboratories, Houston, Texas) are somewhat specific, they lack sensitivity (Richardson and Kokki 1999). Furthermore, the detection of Candida sp. antibodies may also be attributed to a mucocutanous infection and is no proof for an invasion, and antibody production may fail to appear in case of immunocompromised patients. b-D-glucan, a component of the fungal cell wall, has become a diagnostic parameter for IFIs (Alam et al. 2007; Odabasi et al. 2004). The available assays use saccharide-specific antibodies or limulus amebocyte lysates (LAL) of the horseshoe crab, which is used for years for endotoxin detection to measure the non-Candida-specific 1,3-b-D-glucan level. The concentration of 1,3-b-D-glucan in plasma of patients with IFI was markedly increased. However, the LAL-based assays from Japan (e.g., Fungitec-G1; Lü et al. 2007) or USA (e.g., Glucatell1), are still not validated for the sole diagnosis of IFIs. They cannot be applied to differentiate aspergillosis from yeast fungus infection, are soonest to be used 3 days postoperatively, and deliver different assay-specific results. 15 New Approaches in Fungal DNA Preparation from Whole Blood 321 D-arabinitol, a further putative target metabolite and exclusive fungal metabolic product is not produced by C. glabrata and C. krusei and is therefore of restricted diagnostic relevance for fungi in general (Walsh et al. 1995). Fungal serum mannan acts as an early antigen marker for candidaemia (Fujita et al. 2006). The Platelia1 Candida Antigen EIA (Bio-Rad; sensitivity 42%, specificity 98%) may be tested negative in patients with anti-mannan antibodies. Therefore, determination of the anti-mannan antibody concentration is recommended. The sensitivity of the combined assay increases to 76%, the specificity to 93% (Sendid et al. 2003). Fungal serum mannan and at least fungal enolase tests (Yeo and Wong 2002) might at least be promising for the early detection of invasive candidiasis, but show short serum half-times and the latter affords complex measurement techniques. In general, antibodies directed against fungal targets are of epidemiological value but antigenic extracts are not yet standardized. The proteins are first seen in a late stadium within the course of the disease and are in general of limited use in immunosuppressed patients. Serological test may in general contribute to IFI diagnostics but have to be assessed by further cultural and clinical data or combined even with current molecular diagnostics (Yeo and Wong 2002). While clinical manifestations of inflammations are elusive, some biochemical parameters indicate early stages of fungal or bacterial infections. The presence of three biomarkers was proven in systemic infections: C-reactive protein (CRP), procalcitonin (PCT), and D-dimer. None of them, however, functions as single independent criterion for sepsis (De La Rosa et al. 2008). CRP, an acute-phase protein that plays a significant role within the complement pathway, binds to several polysaccharides present in all classes of sepsis causative pathogens. Secretion of CRP starts within several hours of the stimulus peaking between 36 and 50 h. After disappearance of the stimulus, CRP falls rapidly with a half-life of 19 h, but it can remain elevated, even for very long periods, if the underlying cause of the elevation persists. Only interventions affecting the inflammatory process responsible for the acute phase reaction can change the CRP level. Changes may be very helpful in diagnosis as well as in monitoring response to therapy, as CRP levels are only determined by the rate of synthesis (Povoa 2002). PCT is a 116-amino acid prohormone proven to be a useful marker in sepsis and sepsis-like conditions (e.g., severe burns and mechanical trauma), and also in some infections of nonbacterial causation as systemic fungal infection (Becker et al. 2004). Serum levels of PCT are frequently increased in sepsis patients, sometimes attaining levels several thousand-fold normal, and these high levels often persist for a long period of time. Moreover, the levels often correlate positively with the severity of the condition and mortality (Pettila et al. 2002; Meisner et al. 1999). A good correlation was found between the serum PCT level and the Sepsis-Related Organ Failure Assessment (SOFA) score, although no correlation was found between the latter and the CRP level (Endo et al. 2008). It has been argued that the PCT serum level may be useful in aiding the diagnosis of sepsis and the discrimination between the specific phases of the disease. Activation of the coagulation cascade is a common and early phenomenon in the development of sepsis, and this fact supports the use of anticoagulant treatments 322 R.P.H. Schmitz et al. as potentially useful interventions (Levi and Ten Cate 1999). Anticoagulation releases degradation products containing D-dimers whereas a finding of more than 500 ng of D-dimer per milliliter is considered abnormal, and such levels are present in virtually all patients with sepsis (Opal and Esmon 2003). However, none of these potential biochemical test parameters which gain information on the status of inflammation and altered along with the development of sepsis, especially parameters regarding the host-response on invasive infections, pathogen cell components, or at least metabolic by-products, do permit access to a directed therapy of the disease. This exclusively demands the identification of the causative agent and putative antibiotic resistances which it exhibits. Various commercial cultural systems that are able to identify sepsis causative pathogens within 4–72 h have been developed. Although correct identification of clinically relevant yeast strains can be achieved with these systems, incomplete or incorrect identifications may occur when certain new and emerging yeast strains are tested (Espinel-Ingroff et al. 1998). The gold standard technique blood culture, which is the routine method in clinical microbiology laboratories to demonstrate the presence of pathogens in patients suspected of systemic infections, exhibits some drawbacks due to the patient’s antibiotic treatment prior to sample withdrawal, a low abundance of causative agents in the blood samples, and frequently noncultivable organisms. The blood culture remains negative in 80–90% of all sepsis incidents and takes a period of usually 24–72 h to obtain results, whereas a sample can be reliably declared negative within up to 7 days. The culture-based phenotypic identification of Candida species from clinical materials, for example, requires at least another day to obtain pure cultures. The results then become the basis for further microbial diagnostics, e.g., species differentiation, biochemical typing, and/or generation of antifungal or antibacterial susceptibility profiles, which are also laborious and time-consuming although the methods are in part automated. 15.3 Fate of Antimycosis The time-to-result of blood cultures is too long to initiate an effective sepsis therapy. Therefore, broad spectrum antibiotics are given simultaneously or prior to blood withdrawal without trusted microbiological findings which enhances the broadening of multi-resistant organisms and downsizes the efforts of later taken blood cultures (Chastre 2008; de Kraker and van de Sande-Bruinsma 2007; Weinstein 2003). Therapy is usually administered indirectly early in the clinical course of blood stream infections (BSI) and prior to reporting of positive blood culture results. With respect to antimicrobial management, the most important information provided by the clinical microbiology laboratory appeared to be the reporting of positive blood culture and Gram stain results. Antimicrobial susceptibility testing data did not appear to have a comparatively important impact on antimicrobial management among patients with BSI (Munson et al. 2003). 15 New Approaches in Fungal DNA Preparation from Whole Blood 323 In general, considerable adverse effects have to be faced when antimycotics are given (e.g., Amphotericin B). Crucial factors for the prognosis of patients with systemic mycoses therefore are again start and accuracy of the antimycotic. Additionally, several human fungal pathogens are characterized by high rates of intrinsic resistance. Therefore, identification of fungal pathogens to the species level rather than antifungal drug susceptibility testing is presently the most important step in selection of adequate antifungal agents (Rex and Pfaller 2002). As expected, IFIs have become an important cause of morbidity and mortality among immunocompromised patients, many of whom are undergoing long-term treatment with antifungal agents. Although there is little evidence of emerging resistance in e.g., Candida albicans (Pfaller et al. 2005), long-term treatment may lead to an increase of non-C. albicans strains (Marr 2004), Aspergillus terreus, and Zygomycetes infections (Wingard 2005; Kauffman 2004). Some Candida species are problematic in this respect notwithstanding: C. glabrata, C. rugosa, and C. guilliermondi display low susceptibilities to fluconazole, an otherwise effective, inexpensive broad spectrum antibiotic with excellent penetration and oral absorption properties (this resistance may also come along with voriconazol resistance), and C. krusei, which is innately resistant like Aspergillus spp. and other molds (Pfaller et al. 2005; Pappas et al. 2003). In general, candidiasis and aspergillosis are the most common IFIs in patients receiving immunosuppressive treatment, e.g., chemotherapy for cancer or organ transplantation, or in immunodeficient patients, such as patients with AIDS (Ellis et al. 2001; Dasbach et al. 2000; Coleman et al. 1998; Barnes et al. 1996), and in addition to the increasing incidence of IFIs, the number of fungal species which must be considered as potential fungal pathogens has also increased during the last few decades. 15.4 Pathogen Detection by Nucleic Acids Amplification Techniques (NAT) Nucleic acid amplification techniques (NAT, e.g., polymerase chain reaction, PCR) allow a more rapid target and resistance detection within several hours, even from whole blood, compared to culture-based methods. Free fungal and bacterial DNAs, as well as DNA from adherent, phagocytosed, and free intact and nonintact pathogens, are detected while blood cultures contribute only in the presence of viable and metabolic active cells. However, the high sensitivity is decreased by factors such as high fractions of salts, hemin, and other blood ingredients, most of which can be effectively removed by affinity chromatography steps during sample preparation. Foremost human bulk DNA, co-isolated with microbial pathogen nucleic acids from the addressed sample material (e.g., whole blood), is additionally the cause for a minute pathogen to human DNA ratio, increased cross-reactivities with primers, and significantly reduced overall analytical assay sensitivities (Handschur et al. 2009). 324 R.P.H. Schmitz et al. In recent years, numerous nucleic acid-based methods have been developed to improve the diagnosis of mycotic infections and the identification of pathogenic fungi (Mikami 2008; Gottfredsson et al. 1998; Reiss et al. 1998; Walsh and Chanock 1998). Prompt and accurate detection and identification of yeast species are important due to the fact that the virulence of Candida isolates for example differs according to the species level, with C. albicans being most virulent, followed by C. tropicalis (Wingard 1995). Therefore, protocols have been published for the detection and identification of mainly yeast pathogens, including species or group discrimination with specific (Yong et al. 2008; Flahaut et al. 1998; Kobayashi et al. 1999; Skladny et al. 1999; Mannarelli and Kurtzman 1998; Reichard et al. 1997) or panfungal (broad-range) PCR primers (Inácio et al. 2008; van Burik et al. 1998). Probes and restriction fragment length polymorphisms have been described in a number of studies to identify unique ribosomal DNA (rDNA) sequences (Evertsson et al. 2000; Kauffman et al. 2000; Loeffler et al. 2000; Martin et al. 2000; Turin et al. 2000; Turenne et al. 1999; Velegraki et al. 1999; Kappe et al. 1998) and as a matter of fact, primers were directed against targets within the rDNA in several approaches (Evertsson et al. 2000; Wahyuningsih et al. 2000; Elie et al. 1998). The identification of PCR products was done by gel or capillary electrophoresis (De Baere et al. 2002; Chen et al. 2000; Turenne et al. 1999), PCR amplicon restriction fragment length polymorphism analysis (Fujita and Hashimoto 2000), single-strand conformational polymorphism (Li and Bai 2007), Southern blot hybridization (microarray) assays with oligonucleotide probes (Spiess et al. 2007; Wiesinger-Mayr et al. 2007; Hebart et al. 2000; Shin et al. 1999; Flahaut et al. 1998; Einsele et al. 1997), and random amplification of polymorphic DNA analysis (Stefan et al. 1997). Furthermore, a number of PCR-enzyme immunoassays (EIAs) has been developed (Badiee et al. 2007; Wellinghausen et al. 2004; Lindsley et al. 2001; Elie et al. 1998; Loeffler et al. 1998; Shin et al. 1997; Fujita et al. 1995). One study showed the applicability of PCR-EIA for the resolution of discrepancies in phenotype-based identification between different institutions (Coignard et al. 2004). Real-time PCR assays have been described for the quantitative detection of either Candida or Aspergillus species in serum (Challier et al. 2004; White et al. 2003; Costa et al. 2001) or other specimen types (White et al. 2004) and Diaz and Fell (2004) applied the Luminex technology for the detection of pathogenic yeasts of the genus Trichosporon. The latter assay combines flow cytometry and nucleotide hybridization via fluorescent beads covalently bound to species-specific capture probes. Upon hybridization, the beads bearing the target amplicons are classified by their spectral addresses with a 635-nm laser. Quantitation of the hybridized biotinylated amplicon is based on fluorescence detection with a 532-nm laser. Although most of these published PCR-based methods have been useful for the identification of fungal species, they either identify only one species at a time or require a probe hybridization procedure that incurs time and expense. An economically more efficient approach would be the application of protocols capable of identifying broad panels of relevant species in a highly parallel fashion (PalkaSantini et al. 2009; Louie et al. 2008). Additionally, the protocols exclude an optimized procedure for the lysis of fungal cells within the sample material, e.g., 15 New Approaches in Fungal DNA Preparation from Whole Blood 325 whole blood or serum, which is a crucial step in nucleic acid detection, and mostly relies on in-house techniques for cell disruption. Beyond any attempts for standardization, the extraction methods for pathogen DNA, i.e., the combination of cell lysis and DNA isolation protocols, produce markedly differing yields of fungal (and bacterial) DNA and thus can significantly affect the results of NAT methods for the respective species (Metwally et al. 2008; Fredricks et al. 2005). Anyway, mechanical impact has been regarded as the most efficient lysis method for fungal cells (Wong et al. 2007; Müller et al. 1998). Further studies described enzymatic/thermal lysis protocols as superior to mechanical impact (Lugert et al. 2006), but the experience should be argued to be mainly attributed to the mechanical device used and its effective power, lysis matrix, sample to headspace volume ratio within the lysis tubes, and the protocol (device settings) applied. It has to be considered that the lastly cited studies on cell lysis were done with pure fungal cells devoid of human cell material, which significantly affects the lysis of the target cells, but the results of Fredricks et al. (2005), which likewise support mechanical lysis, were obtained from spiked clinical samples, a technique which should basically be recommended for assay development. However, the mechanical devices of the last generation (e.g., FastPrep1-24, MP Biomedicals, Solon, OH, USA), which disrupt cells in periods of several seconds to few minutes while executing figure-8 vertical, angular motions, seem to guarantee for the first time acceptable efficiencies for standardized disruptions of designated rigid pathogen cells within high-volume (i.e., 2 mL) clinical samples. The cells are ground mostly between glass or ceramic bead sand mixtures of particle sizes between 0.1 and 2.5 mm in diameter. Higher head-space volumes over the sample/ bead-matrix suspension support the lysis efficiency, an effect which is obviously noticed in tubes of 15 mL total volume (suspension to head-space volume 2:1). Although mechanical lysis is described as an inexpensive technique, the rotating equipment which ensures sufficient efficiency, has to be regarded as initially quite cost-intensive. Additionally, a DNA extraction method is indispensable for a PCRbased assay and both manual and automated protocols are forthcoming (Loeffler et al. 2002) – the latter often high priced. A recently launched diagnostic PCR-linked tool bases on a selective lysis of (human) blood cells by chaotropic buffers and quantitative degradation of human nucleic acids via chaotroph-resistant DNase (Horz et al. 2008). The enzyme is subsequently heat-inactivated and bacterial pathogen cells are lyzed in a second step to gain their genomic DNA for NAT detection purposes. It has to be questioned if this initially smartly appearing two-step lysis procedure is actually a drawback of the analytical approach, since sublethally affected pathogen cells (e.g., due to antibiotic treatment or impact of the immune system) and Gram-negative thin-walled cells may be disrupted within the first lysis step and lose their genomic (target) DNA (Horz et al. 2008). However, the tool may preferably be suited for rigid cell types, but its applicability for the detection of yeast cells, which has been declared to be covered with the assay panel, has not yet been proven in clinical studies. A new multiplex PCR-based assay (VYOO1, SIRS-Lab GmbH, Jena) was launched recently and compensates the above mentioned drawbacks by combining 326 R.P.H. Schmitz et al. an efficient bead-based mechanical lysis protocol with an extended multiplex PCR detection step (Sachse et al. 2009; Horz et al. 2008). The launched assay includes a unique pretreatment tool which specifically concentrates fungal and bacterial DNA by affinity chromatography and removes human background DNA, the cause for a minute fungal plus bacterial to human DNA ratio within the addressed sample material (e.g., EDTA whole blood). A truncated DNA-binding protein recognizes unmethylated XpYpCpGpXpY motifs within the DNA which are significantly less frequent in humans than in yeast (and bacterial) DNA (Pinarbasi et al. 1996; Wilkinson et al. 1995). An earlier study described different methylation grades between the mycelia and yeast forms of C. albicans and outlined an overall lower methylation of fungal DNA compared to DNA from higher eukaryotes (Russell et al. 1987). However, the enrichment of pathogen DNA significantly increases the sensitivity of the chosen downstream detection method by at least one order of magnitude (Sachse et al. 2009). The standard (as compared to real-time PCR/qPCR assays) multiplex PCR detects an optimized panel of six fungal (A. fumigatus, C. albicans, C. glabrata, C. krusei, C. parapsilosis, and C. tropicalis) (and 34 bacterial) species that cause life-threatening infections and comprise 99% of sepsis-causative pathogens (as confirmed by the frequencies of pathogens via meta-analysis of 11 clinical studies, see above legend to Fig. 15.1), as well as a set of important antibiotic resistances (e.g., methicillin, vancomycin, b-lactamase). The test has been designed for the examination of the generated amplicons by gel-electrophoresis or hybridization methodologies. The latter is demanded for increased sensitivities and result verification. However, using amplicon-specific DNA length markers, the amplicon identification works straightforward and allows for results within 8 h – an important benefit in sepsis diagnostics compared to assays which rely on preculturing of viable cells (i.e., blood culture) (Lau et al. 2008) (Fig. 15.3). The crucial step in pathogen detection from whole blood is the low abundance of pathogen cells within the circuit. A high percentage of bacteraemia gained positive blood cultures from blood samples possessing less than 10 cfu of pathogen cells per Fig. 15.3 Time course of the detection of fungal causative pathogens with VYOO1 15 New Approaches in Fungal DNA Preparation from Whole Blood 327 mL (Kreger et al. 1980). Sine qua non for a new NAT-based assay is therefore a sensitivity which enters this analytical range to compete with the standard techniques in the lab and with the medical tradition which has to be persuaded for the effectiveness of nucleic acid based assay systems. C. albicans, for example, was detected in EDTA whole blood after spiking, using the new multiplex PCR tool and gelelectrophoretic amplicon detection with a sensitivity of <20 cfu/mL, a level which was attained previously only by qPCR-based techniques (Lehmann et al. 2008) requiring high-priced analytical devices and confined by restricted numbers of detectable targets due to the available wavelength-specific fluorescence dyes (e. g., 6-carboxy-fluorescein/FAM) and/or fluorescence channels of the qPCR device used. Their number is in fact growing with the state of the art, but to achieve a still limited but increased multiplex applicability fluorescence detection (e.g., via dual fluorescence resonance energy transfer probes targeting species-specific internal transcribed spacer (ITS) regions of PCR amplicons (Dunyach et al. 2008; Mancini et al. 2008)), is combined with error-prone melting point analyzes and done with a high workload (Schrenzel 2007). The progress of the new standard multiplex procedure has in part to be attributed to the high sample volume of 5 mL whole blood which can be applied, regardless of the benefits of the usage of 15 mL centrifugation vials within the lysis step (see above). The importance of the blood volume cultured or processed in NAT-based methods has been discussed long ago (Arpi et al. 1989); however, the implications of reduced sample volumes has been accepted for the qPCR evaluation due to the always aspired miniaturization of assay layouts. The innovative part of the new multiplex PCR assay, however, remains the protein-based pathogen-DNA affinity chromatography step which is additionally sold as an assay-independent pretreatment tool (LOOXSTER1), which may substantially enhance even home-brew NAT-based assays, and the unmatched ultrahigh multiplex assay for the simultaneous detection of fungal (and bacterial) pathogens within one working day. 15.5 Conclusion and Future Line of Research In recent years, a broad spectrum of PCR-based methods for the detection of sepsis causative fungal pathogens has been developed and some of them might be suitable to support and expedite clinical findings and strengthen a directed and restricted antibiotic therapy. A standardization, however, is not yet in sight. At present, only sparse information is available on potential cost benefits for application of molecular diagnosis versus conventional detection of bloodstream infections (Falagas and Panayotis 2008) but it has to be assumed, that the high initial and running costs will be outweighed by the increased assay sensitivity and reduced time-to-result. The assays are in part compromised by their exquisite sensitivities responsible for nucleic acid trace detection already present in associated consumables or introduced via applying routine sample withdrawal techniques causing false-positive results. 328 R.P.H. Schmitz et al. Origin and clinical significance of those “false-positive” samples are often ambiguous and might belong to yet unbeknown host-pathogen interactions (Schrenzel 2007). Positive NAT-related findings therefore have to be correlated with other clinical observations and diagnostic tests by clinicians and should not be the sole reason for any therapeutical consequences. It should be remembered that three decades ago culture-positive bacteraemia was reported after tooth brushing (Berger et al. 1974), which indicates the range of risk factors that obtain false-positives and the difficulties in distinguishing them from true-positives, a problem which might also be assigned to fungaemia in general. The usage of broad-range primers should be consciously balanced in favor of particular species detections. The amplification method itself, e.g., loop-mediated isothermal amplification (LAMP), ligase chain reaction, nucleic acid sequence-based amplification (NASBA), self-sustained sequence replication (3SR), strand displacement amplification (SDA), transcription-mediated amplification (TMA), or cycling probe technology (CPT), as listed by Schrenzel (2007), may in part significantly alter and improve the sensitivity of the invented verification procedure – the NAT itself will only be as suitable as the associated total DNA release and isolation protocols are – and of course, the approach of getting rid of bulk DNA and polymerase inhibitors. What is required are combined multitarget assay systems with proven clinical utility and that offer high negative predictive values with notwithstanding high analytical sensitivities and low detection thresholds. Further approaches in test development should focus on standardization and ongoing shortening of the clinical course, e.g., via automation of test flows. Coated magnetic bead particles, e.g., for covalent binding of proteins and antibodies (e.g., for affinity chromatography), capture of biotinylated biomolecules (e.g., for hybridization techniques), DNA separation, or immunological applications (e.g., enzyme-linked immunosorbent assays, ELISA), are already used in many applications, since centrifugation is not required due to the separation of the beads, sized from 50 nm to approximately 3 mm in diameter, from aqueous phases with magnets. Time-consuming DNA-sedimentations and redilutions are circumvented. The combination of multiplex PCR with microarrays on chips in part placed at the bottom of reaction vials (e.g., ArrayTube1 system, Clondiag, Jena, Germany) represents a currently tedious and laborious but forward-looking development. A new gold standard for the detection of fungal pathogens, however, not yet found, will for sure be nucleic acid based as announced by the current diagnostic achievements. Acknowledgments The work was supported by grants from the Thuringian Ministry of Economy, Technology and Labour (TMWTA 2007 FE 0255). References Alam FF, Mustafa AS, Khan ZU (2007) Comparative evaluation of (1, 3)-b-D-glucan, mannan and anti-mannan antibodies, and Candida species-specific snPCR in patients with candidaemia. BMC Infect Dis 7:103 15 New Approaches in Fungal DNA Preparation from Whole Blood 329 Alvarez-Lerma F (1996) Modification of empiric antibiotic treatment in patients with pneumonia acquired in the intensive care unit. ICU-Acquired Pneumonia Study Group. Intensive Care Med 22:387–394 Angus DC, Wax RS (2001) Epidemiology of sepsis: an update. Crit Care Med 29:S109–S116 Arpi M, Bentzon MW, Jensen J, Frederiksen W (1989) Importance of blood volume cultures in the detection of bacteraemia. Eur J Clin Microbiol Infect Dis 8:838–842 Ascioglu S, Rex JH, de Pauw B, Bennett JE, Bille J, Crokaert F, Denning DW, Donnelly JP, Edwards JE, Erjavec Z, Fiere D, Lortholary O, Maertens J, Meis JF, Patterson TF, Ritter J, Selleslag D, Shah PM, Stevens DA, Walsh TJ (2002) Defining opportunistic invasive fungal infections in immunocompromised patients with cancer and hematopoietic stem cell transplants: an international consensus. Invasive Fungal Infections Cooperative Group of the European Organization for Research and Treatment of Cancer. Mycoses Study Group of the National Institute of Allergy and Infectious Diseases. Clin Infect Dis 34:7–14 Badiee P, Kordbacheh P, Alborzi A, Malekhoseini S, Zeini F, Mirhendi H, Mahmoodi M (2007) Prospective screening in liver transplant recipients by panfungal PCR-ELISA for early diagnosis of invasive fungal infections. Liver Transpl 13:1011–1016 Barnes RA, Denning DW, Evans EGV, Hay RJ, Kibbler CC, Prentice AG, Richardson MD, Roberts MM, Rogers TR, Speller DC, Warnock DW, Warren RE (1996) Fungal infections: a survey of laboratory services for diagnosis and treatment. Commun Dis Rep Rev 6:R69–R75 Bauer M, Brunkhorst FM, Welte T, Gerlach H, Reinhart K (2006) Sepsis – aktuelle Aspekte zur Pathophysiologie, Diagnostik und Therapie. Anaesthesist 55:835–845 Becker KL, Nylen ES, White JC, Muller B, Snider RH Jr (2004) Clinical review 167: procalcitonin and the calcitonin gene family of peptides in inflammation, infection, and sepsis: a journey from calcitonin back to its precursors. J Clin Endocrinol Metab 89:1512–1525 Berger SA, Weitzman S, Edberg SC, Casey JI (1974) Bacteraemia after the use of an oral irrigation device. A controlled study in subjects with normal-appearing gingiva: comparison with use of a toothbrush. Ann Intern Med 80:510–511 Bodmann K-F, Vogel F (2001) Antimikrobielle Therapie der Sepsis. Chemother J 10:43–56 Challier S, Boyer S, Abachin E, Berche P (2004) Development of a serum-based TaqMan realtime PCR assay for diagnosis of invasive aspergillosis. J Clin Microbiol 42:844–846 Charles PE, Doise JM, Quenot JP, Aube H, Dalle F, Chavanet P, Milesi N, Aho LS, Portier H, Blettery B (2003) Candidaemia in critically ill patients: difference of outcome between medical and surgical patients. Intensive Care Med 29:2162–2169 Chastre J (2008) Evolving problems with resistant pathogens. Clin Microbiol Infect 14:3–14 Chen YC, Eisner JD, Kattar MM, Rassoulian-Barrett SL, Lafe K, Yarfitz SL, Limaye AP, Cookson BT (2000) Identification of medically important yeasts using PCR-based detection of DNA sequence polymorphisms in the internal transcribed spacer 2 region of the rRNA genes. J Clin Microbiol 38:2302–2310 Cockerill FR 3rd, Hughes JG, Vetter EA, Mueller RA, Weaver AL, Ilstrup DM, Rosenblatt JE, Wilson WR (1997) Analysis of 281, 797 consecutive blood cultures performed over an 8-year period: trends in micro-organisms isolated and the value of anaerobic culture of blood. Clin Infect Dis 24:403–418 Coignard C, Hurst SF, Benjamin LE, Brandt ME, Warnock DW, Morrison CJ (2004) Resolution of discrepant results for Candida species identification by using DNA probes. J Clin Microbiol 42:858–861 Coleman DC, Rinaldi MG, Haynes KA, Rex JH, Summerbell RC, Anaissie EJ, Li A, Sullivan DJ (1998) Importance of Candida species other than Candida albicans as opportunistic pathogens. Med Mycol 36:156–165 Costa C, Vidaud D, Olivi M, Bart-Delabesse E, Vidaud M, Bretagne S (2001) Development of two real-time quantitative TaqMan PCR assays to detect circulating Aspergillus fumigatus DNA in serum. J Microbiol Methods 44:263–269 Dasbach EJ, Davies GM, Teutsch SM (2000) Burden of aspergillosis-related hospitalizations in the United States. Clin Infect Dis 31:1524–1528 330 R.P.H. Schmitz et al. De Baere T, Claeys G, Swinne D, Verschraegen G, Muylaert A, Massonet C, Vaneechoutte M (2002) Identification of cultured isolates of clinically important yeast species using fluorescent fragment length analysis of the amplified internally transcribed rRNA spacer 2 region (ITS2). BMC Microbiol 2:21 de Kraker M, van de Sande-Bruinsma N (2007) Trends in antimicrobial resistance in Europe: update of EARSS results. Euro Surveill 12:E070315.3 De La Rosa GD, Valencia ML, Arango CM, Gomez CI, Garcia A, Ospina S, Osorno S, Henao A, Jaimes FA (2008) Toward an operative diagnosis in sepsis: a latent class approach. BMC Infect Dis 8:18 Diaz MR, Fell JW (2004) High-throughput detection of pathogenic yeasts of the genus Trichosporon. J Clin Microbiol 42:3696–3706 Dunyach C, Bertout S, Phelipeau C, Drakulovski P, Reynes J, Mallié M (2008) Detection and identification of Candida spp. in human serum by LightCycler real-time polymerase chain reaction. Diagn Microbiol Infect Dis 60:263–271 Duthie R, Denning DW (1995) Aspergillus fungaemia: report of two cases and review. Clin Infect Dis 20:598–605 Einsele H, Hebart H, Roller G, Loffler J, Rothenhofer I, Muller CA, Bowden RA, van Burik J, Engelhard D, Kanz L, Schumacher U (1997) Detection and identification of fungal pathogens in blood by using molecular probes. J Clin Microbiol 35:1353–1360 Elie CM, Lott TJ, Reiss E, Morrison CJ (1998) Rapid identification of Candida species with species-specific DNA probes. J Clin Microbiol 36:3260–3265 Ellis ME, Al-Abdely H, Sandridge A, Greer W, Ventura W (2001) Fungal endocarditis: evidence in the world literature, 1965–1995. Clin Infect Dis 32:50–62 Endo S, Aikawa N, Fujishima S, Sekine I, Kogawa K, Yamamoto Y, Kushimoto S, Yukioka H, Kato N, Totsuka K, Kikuchi K, Ikeda T, Ikeda K, Yamada H, Harada K, Satomura S (2008) Usefulness of procalcitonin serum level for the discrimination of severe sepsis from sepsis: a multicenter prospective study. J Infect Chemother 14:244–249 Espinel-Ingroff A, Stockman L, Roberts G, Pincus D, Pollack J, Marler J (1998) Comparison of RapID Yeast Plus system with API 20C system for identification of common, new, and emerging yeast pathogens. J Clin Microbiol 36:883–886 Evertsson U, Monstein HJ, Johansson AG (2000) Detection and identification of fungi in blood using broad-range 28S rDNA PCR amplification and species-specific hybridisation. Acta Pathol Microbiol Immunol Scand 108:385–392 Falagas ME, Panayotis TT (2008) From acute renal failure to acute kidney injury: emerging concepts. Crit Care Med 36:1641–1642 Fine JM, Fine MJ, Galusha D, Petrillo M, Mechan TP (2002) Patient and hospital characteristics associated with recommended processes of care for elderly patients hospitalized with pneumonia. Arch Intern Med 162:827–833 Flahaut M, Sanglard D, Monod M, Bille J, Rossier M (1998) Rapid detection of Candida albicans in clinical samples by DNA amplification of common regions from C. albicans-secreted aspartic proteinase genes. J Clin Microbiol 36:395–401 Focht J, Adam D (2004) Antibakterielle Aktivität verschiedener Antibiotika. Klinikarzt 33:35–38 Fraser VJ, Jones M, Dunkel J, Storfer S, Medoff G, Dunagan WC (1992) Candidaemia in a tertiary care hospital: epidemiology, risk factors, and predictors of mortality. Clin Infect Dis 15:414–421 Fredricks DN, Smith C, Meier A (2005) Comparison of six DNA extraction methods for recovery of fungal DNA as assessed by quantitative PCR. J Clin Microbiol 43:5122–5128 Fridkin SK, Gaynes RP (1999) Antimicrobial resistance in intensive care units. Clin Chest Med 20:303–316 Fujita S, Hashimoto T (2000) DNA fingerprinting patterns of Candida species using HinfI endonuclease. Int J Syst Evol Microbiol 50:1381–1380 Fujita S, Lasker BA, Lott TJ, Reiss E, Morrison CJ (1995) Microtitration plate enzyme immunoassay to detect PCR-amplified DNA from Candida species in blood. J Clin Microbiol 33:962–967 15 New Approaches in Fungal DNA Preparation from Whole Blood 331 Fujita S, Takamura T, Nagahara M, Hashimoto T (2006) Evaluation of a newly developed downflow immunoassay for detection of serum mannan antigens in patients with candidaemia. J Med Microbiol 55:537–543 Gamis A, Gudnason T, Giebink G, Ramsay N (1991) Disseminated infection with Fusarium in recipients of bone marrow transplants. Rev Infect Dis 13:1077–1088 Garnacho-Montero J, Garcia-Garmendia JL, Barrero-Almodovar A, Jimenez-Jimenez FJ, Perez-Paredes C, Ortiz-Leyba C (2003) Impact of adequate empirical antibiotic therapy on the outcome of patients admitted to the intensive care unit with sepsis. Crit Care Med 31:2742–2751 Geerdes HF, Ziegler D, Lode H, Hund M, Loehr A, Fangmann W, Wagner J (1992) Septicemia in 980 patients at a university hospital in Berlin: prospective studies during 4 selected years between 1979 and 1989. Clin Infect Dis 15:991–1002 Geffers C, Zuschneid I, Sohr D, Rüden H, Gastmeier P (2004) Erreger nosokomialer Infektionen auf Intensivstationen: Daten des Krankenhaus-Infektions-Surveillance-Systems (KISS) aus 274 Intensivstationen. Anästhesiol Intensivmed Notfallmed Schmerzther 39:15–19 Gottfredsson M, Cox GM, Perfect JR (1998) Molecular methods for epidemiological and diagnostic studies of fungal infections. Pathology 30:405–418 Gudlaugsson O, Gillespie S, Lee K, Van de Berg J, Hu J, Messer S, Herwaldt L, Pfaller M, Diekema D (2003) Attributable mortality of nosocomial candidaemia, revisited. Clin Infect Dis 37:1172–1177 Handschur M, Karlic H, Hertel C, Pfeilstöcker M, Haslberger AG (2009) Preanalytic removal of human DNA eliminates false signals in general 16S rDNA PCR monitoring of bacterial pathogens in blood. Comp Immunol Microbiol Infect Dis 32:207–219 Hebart H, Loffler J, Meisner C, Serey F, Schmidt D, Bohme A, Martin H, Engel A, Bunje D, Kern WV, Schumacher U, Kanz L, Einsele H (2000) Early detection of Aspergillus infection after allogeneic stem cell transplantation by polymerase chain reaction screening. J Infect Dis 181:1713–1719 Horz HP, Scheer S, Huenger F, Vianna ME, Conrads G (2008) Selective isolation of bacterial DNA from human clinical specimens. J Microbiol Methods 72:98–102 Ibrahim EH, Sherman G, Ward S, Fraser VJ, Kollef MH (2000) The influence of inadequate antimicrobial treatment of bloodstream infections on patient outcomes in the ICU setting. Chest 118:146–155 Inácio J, Flores O, Spencer-Martins I (2008) Efficient identification of clinically relevant Candida yeast species by use of an assay combining panfungal loop-mediated isothermal DNA amplification with hybridization to species-specific oligonucleotide probes. J Clin Microbiol 46:713–720 Iregui M, Ward S, Sherman G, Fraser VJ, Kollef MH (2002) Clinical importance of delays in the initiation of appropriate antibiotic treatment for ventilator-associated pneumonia. Chest 122:262–268 Jones ME, Draghi DC, Thornsberry C, Karlowsky JA, Sahm DF, Wenzel RP (2004) Emerging resistance among bacterial pathogens in the intensive care unit – a European and North American surveillance study (2991–2002). Ann Clin Microbiol Antimicrob 3:1–11 Kappe R, Okeke CN, Fauser C, Maiwald M, Sonntag H-G (1998) Molecular probes for the detection of pathogenic fungi in the presence of human tissue. J Med Microbiol 47:811–820 Karlowsky JA, Jones ME, Draghi DC, Thornsberry C, Sahm DF, Volturo GA (2004) Prevalence and antimicrobial susceptibilities of bacteria isolated from blood cultures of hospitalized patients in the United States in 2002. Ann Clin Microbiol Antimicrob 3:1–8 Kauffman CA (2004) Zygomycosis: re-emergence of an old pathogen. Clin Infect Dis 39:588–590 Kauffman CA, Vázquez JA, Sobel JD, Gallis HA, McKinsey DS, Karchmer AW, Sugar AM, Sharkey PK, Wise GJ, Mangi R, Mosher A, Lee JY, Dismukes WE et al (2000) Prospective multicenter surveillance study of funguria in hospitalized patients. Clin Infect Dis 30:14–18 Kobayashi M, Sonobe H, Ikezoe T, Hakoda E, Ohtsuki Y, Taguchi H (1999) In situ detection of Aspergillus 18S ribosomal RNA in invasive pulmonary aspergillosis. Intern Med 38:563–569 332 R.P.H. Schmitz et al. Kreger BE, Craven DE, Carling PC, McCabe WR (1980) Gram-negative bacteraemia III. Reassessment of etiology, epidemiology and ecology in 612 patients. Am J Med 68:332–343 Kübler A, Durek G, Zamirowska A, Duszynska W, Palysinska B, Gaszynski W, Pluta A (2004) Severe sepsis in Poland – results of internet surveillance of 1043 cases. Med Sci Monit 10:635–641 Lau A, Sorrell T, Chen S, Stanley K, Iredell J, Halliday C (2008) Multiplex-tandem PCR: a novel platform for the rapid detection and identification of fungal pathogens from blood culture specimens. J Clin Microbiol 46:3021–3027 Lehmann LE, Hunfeld K-P, Emrich T, Haberhausen G, Wissing H, Hoeft A, Stüber F (2008) A multiplex real-time PCR assay for rapid detection and differentiation of 25 bacterial and fungal pathogens from whole blood samples. Med Microbiol Immunol 197:313–324 Levi M, Ten Cate H (1999) Disseminated intravascular coagulation. N Engl J Med 341:586–592 Li J, Bai FY (2007) Single-strand conformation polymorphism of microsatellite for rapid strain typing of Candida albicans. Med Mycol 45:629–635 Liu M, Healy MD, Dougherty BA, Esposito KM, Maurice TC, Mazzucco CE, Bruccoleri RE, Davison DB, Froscro M, Barrett JF, Wang YK (2006) Conserved fungal genes as potential targets for broad-spectrum antifungal drug discovery. Eukaryotic Cell 5:638–649 Lindsley MD, Hurst SF, Iqbal NJ, Morrison CJ (2001) Rapid identification of dimorphic and yeastlike fungal pathogens using specific DNA probes. J Clin Microbiol 39:3505–3511 Loeffler J, Hebart H, Sepe S, Schumacher U, Klingebiel T, Einsele H (1998) Detection of PCRamplified fungal DNA by using a PCR ELISA system. Med Mycol 36:275–279 Loeffler J, Henke N, Hebart H, Schmidt D, Hagmeyer L, Schumacher U, Einsele H (2000) Quantification of fungal DNA by using fluorescence resonance energy transfer and the light cycler system. J Clin Microbiol 38:586–590 Loeffler J, Schmidt K, Hebart H, Schumacher U, Einsele H (2002) Automated extraction of genomic DNA from medically important yeast species and filamentous fungi by using the MagNA Pure LC system. J Clin Microbiol 40:2240–2243 Louie RF, Tang Z, Albertson TE, Cohen S, Tran NK, Kost GJ (2008) Multiplex polymerase chain reaction detection enhancement of bacteraemia and fungemia. Crit Care Med 36:1487–1492 Lü PH, Zhao BL, Shi Y, Wen YT (2007) The diagnostic value of detecting plasma 1, 3-beta-Dglucan for invasive fungal infections. Zhonghua Jie He He Hu Xi Za Zhi 30:31–34 Lugert R, Schettler C, Gross U (2006) Comparison of different protocols for DNA preparation and PCR for the detection of fungal pathogens in vitro. Mycoses 49:298–304 Luna CM, Vujacich P, Niederman MS, Vay C, Ghepardi C, Matera J, Jolly EC (1997) Impact of BAL data on the therapy and outcome of ventilator-associated pneumonia. Chest 111:676–685 Mancini N, Clerici D, Diotti R, Perotti M, Ghidoli N, De Marco D, Pizzorno B, Emrich T, Burioni R, Ciceri F, Clementi M (2008) Molecular diagnosis of sepsis in neutropenic patients with haematological malignancies. J Med Microbiol 57:601–604 Mannarelli BM, Kurtzman CP (1998) Rapid identification of Candida albicans and other human pathogenic yeasts by using short oligonucleotides in a PCR. J Clin Microbiol 36:1634–1641 Marr KA (2004) Invasive Candida infections: the changing epidemiology. Oncology (Huntington) 18:9–14 Martin GS, Mannino DM (2003) The epidemiology of sepsis in the United States from 1979 through 2000. NEJM 348:1546–1554 Martin C, Roberts D, van der Weide M, Rossau R, Jannes G, Smith T, Maher M (2000) Development of a PCR-based line probe assay for identification of fungal pathogens. J Clin Microbiol 38:3735–3742 Martin GS, Mannino DM, Eaton S, Moss M (2003) The epidemiology of sepsis in the United States from 1979 through 2000. N Engl J Med 348:1546–1554 Meisner M, Tschaikowsky K, Palmaers T, Schmidt J (1999) Comparison of procalcitonin (PCT) and C-reactive protein (CRP) plasma concentrations at different SOFA scores during the course of sepsis and MODS. Crit Care 3:45–50 Metwally L, Fairley DJ, Coyle PV, Hay RJ, Hedderwick S, McCloskey B, O’Neill HJ, Webb CH, Elbaz W, McMullan R (2008) Improving molecular detection of Candida DNA in whole 15 New Approaches in Fungal DNA Preparation from Whole Blood 333 blood: comparison of seven fungal DNA extraction protocols using real-time PCR. J Med Microbiol 57:296–303 Mikami Y (2008) Proposal of new molecular characterization methods in phylogenetic studies and genotypings of pathogenic fungi. Nippon Ishinkin Gakkai Zasshi 49:151–155 Morrell M, Fraser VJ, Kollef MH (2005) Delaying the empiric treatment of Candida bloodstream infection until positive blood culture results are obtained: a potential risk factor for hospital mortality. Antimicrob Agents Chemother 49:3640–3645 Müller FM, Werner KE, Kasai M, Francesconi A, Chanock SJ, Walsh TJ (1998) Rapid extraction of genomic DNA from medically important yeasts and filamentous fungi by high-speed cell disruption. J Clin Microbiol 36:1625–1629 Munson EL, Diekema DJ, Beekmann SE, Chapin KC, Doern GV (2003) Detection and treatment of bloodstream infection: laboratory reporting and antimicrobial management. J Clin Microbiol 41:495–497 Nolla-Salas J, Sitges-Serra A, León-Gil C, Martı́nez-González J, León-Regidor MA, Ibáñez-Lucı́a P, Torres-Rodrı́guez JM (1997) Candidaemia in non-neutropenic critically ill patients: analysis of prognostic factors and assessment of systemic antifungal therapy. Study group of fungal infection in the ICU. Intensive Care Med 23:23–30 Odabasi Z, Mattiuzzi G, Estey E, Kantarjian H, Saeki F, Ridge RJ, Ketchum PA, Finkelman MA, Rex JH, Ostrosky-Zeichner L (2004) Beta-D-glucan as a diagnostic adjunct for invasive fungal infections: validation, cutoff development, and performance in patients with acute myelogenous leukemia and myelodysplastic syndrome. Clin Infect Dis 39:199–205 Opal SM, Esmon CT (2003) Bench-to-bedside review: functional relationships between coagulation and the innate immune response and their respective roles in the pathogenesis of sepsis. Crit Care 7:23–38 Palka-Santini M, Cleven BE, Eichinger L, Krönke M, Krut O (2009) Large scale multiplex PCR improves pathogen detection by DNA microarrays. BMC Microbiol 9:1 Pappas PG, Rex JH, Lee J, Hamill RJ, Larsen RA, Powderly W, Kauffman CA, Hyslop N, Mangino JE, Chapman S, Horowitz HW, Edwards JE, Dismukes WE, for the NIAID Mycoses Study Group (2003) A prospective observational study of candidaemia: epidemiology, therapy, and influences on mortality in hospitalized adult and pediatric patients. Clin Infect Dis 37:634–643 Pettila V, Hynninen M, Takkunen O, Kuusela P, Valtonen M (2002) Predictive value of procalcitonin and interleukin 6 in critically ill patients with suspected sepsis. Intensive Care Med 28:1220–1225 Pfaller MA, Diekema DJ, Rinaldi MG, Barnes R, Hu B, Veselov AV, Tiraboschi N, Nagy E, Gibbs DL, and the Global Antifungal Surveillance Group (2005) Results from the ARTEMIS DISK Global Antifungal Surveillance Study: a 6.5-year analysis of susceptibilities of Candida and other yeast species to fluconazole and voriconazole by standardized disk diffusion testing. J Clin Microbiol 43:5848–5859 Pinarbasi E, Elliott J, Hornby DP (1996) Activation of a yeast pseudo DNA methyltransferase by deletion of a single amino acid. J Mol Biol 257:804–813 Povoa P (2002) C-reactive protein: a valuable marker of sepsis. Intensive Care Med 28:235–243 Reichard U, Margraf S, Hube B, Rüchel R (1997) A method for recovery of Candida albicans DNA from larger blood samples and its detection by polymerase chain reaction on proteinase genes. Mycoses 40:249–253 Reinhart K, Brunkhorst FM, Bone HG, Gerlach H, Gründling M, Kreymann G, Kujath P, Marggraf G, Mayer K, Meier-Hellmann A, Peckelsen C, Putensen C, Stüber F, Quintel M, Ragaller M, Rossaint R, Weiler N, Welte T, Werdan K (2006) Diagnosis and therapy of sepsis. Clin Res Cardiol 95:429–454 Reiss E, Tanaka K, Bruker G, Chazalet V, Coleman DC, Debeaupuis J-P, Hanazawa R, Latgé J-P, Lortholary J, Makimura K, Morrison CJ, Murayama SY, Naoe S, Paris S, Sarfati J, Shibuya K, Sullivan DJ, Uchida K, Yamaguchi H (1998) Molecular diagnosis and epidemiology of fungal infections. Med Mycol 36:249–257 334 R.P.H. Schmitz et al. Rello J, Gallego M, Mariscal D, Soñora R, Vallés J (1997) The value of routine microbial investigation in ventilator-associated pneumonia. Am J Respir Crit Care Med 156:196–200 Rex JH, Pfaller MA (2002) Has antifungal susceptibility testing come of age? Clin Infect Dis 35:982–989 Richardson MD, Kokki MH (1999) New perspectives in the diagnosis of systemic fungal infection. Ann Med 31:327–335 Russell PJ, Welsch JA, Rachlin EM, McCloskey JA (1987) Different levels of DNA methylation in yeast and mycelial forms of Candida albicans. J Bacteriol 169:4393–4395 Sachse S, Straube E, Lehmann M, Bauer M, Rußwurm S, Schmidt K-H (2009) Truncated hCGBP improves NAT-based detection of bacterial species in human samples. J Clin Microbiol 47:1050–1057 Schmidt U, Pfaffenbach B, Quabeck K, Donhuijsen K (1991) Fungal infections after bone marrow transplantation – an autopsy study. Mycoses 34:S33–35 Schrenzel J (2007) Clinical relevance of new diagnostic methods for bloodstream infections. Int J Antimicrob Agents 30:S2–6 Sendid B, Caillot D, Baccouch-Humbert B, Klingspor L, Grandjean M, Bonnin A, Poulain D (2003) Contribution of the Platelia Candida-specific antibody and antigen tests to early diagnosis of systemic Candida tropicalis infection in neutropenic adults. J Clin Microbiol 41:4551–4558 Shin JH, Nolte FS, Morrison CJ (1997) Rapid identification of Candida species in blood cultures by a clinically useful PCR method. J Clin Microbiol 35:1454–1459 Shin JH, Nolte FS, Holloway BP, Morrison CJ (1999) Rapid identification of up to three Candida species in a single reaction tube by a 50 exonuclease assay using fluorescent DNA probes. J Clin Microbiol 37:165–170 Skladny H, Buchheidt D, Baust C, Krieg-Schneider F, Seifarth W, Leib-Mosch C, Hehlmann R (1999) Specific detection of Aspergillus species in blood and bronchoalveolar lavage samples of immunocompromised patients by two-step PCR. J Clin Microbiol 37:3865–3871 Slade E, Tamber PS, Vincent JL (2003) The surviving sepsis campaign: raising awareness to reduce mortality. Crit Care 7:1–2 Spiess B, Seifarth W, Hummel M, Frank O, Fabarius A, Zheng C, Mörz H, Hehlmann R, Buchheidt D (2007) DNA microarray-based detection and identification of fungal pathogens in clinical samples from neutropenic patients. J Clin Microbiol 45:3743–3753 Stefan P, Vazquez JA, Boikov D, Xu C, Sobel JD, Akins RA (1997) Identification of Candida species by randomly amplified polymorphic DNA fingerprinting of colony lysates. J Clin Microbiol 35:2031–2039 Styers D, Sheehan DJ, Hogan P, Sahm DF (2006) Laboratory-based surveillance of current antimicrobial resistance patterns and trends among Staphylococcus aureus: 2005 status in the United States. Ann Clin Microbiol Antimicrob 5:2 Turenne CY, Sanche SE, Hoban DJ, Karlowsky JA, Kabani AM (1999) Rapid identification of fungi by using the ITS2 genetic region and an automated fluorescent capillary electrophoresis system. J Clin Microbiol 37:1846–1851 Turin L, Riva F, Galbiati G, Cainelli T (2000) Fast, simple and highly sensitive double-rounded polymerase chain reaction assay to detect medically relevant fungi in dermatological specimens. Eur J Clin Investig 30:511–518 Vallés J, Rello J, Ochagavı́a A, Garnacho J, Alcali MA (2003) Community-acquired blood-stream infection in critically ill adult patients: impact of shock and inappropriate antibiotic therapy on survival. Chest 123:1615–1624 van Burik J-A, Myerson D, Schreckhise RW, Bowden RA (1998) Panfungal PCR assay for detection of fungal infection in human blood specimens. J Clin Microbiol 36:1169–1175 Velegraki A, Kambouris ME, Skiniotis G, Savala M, Mitroussia-Ziouva A, Legakis NJ (1999) Identification of medically significant fungal genera by polymerase chain reaction followed by restriction enzyme analysis. FEMS Immunol Med Microbiol 23:303–312 Vincent JL, Anaissie E, Bruining H, Demajo W, El-Ebiary M, Haber J, Hiramatsu Y, Nitenberg G, Nystrom PO, Pittet D, Rogers T, Sandven P, Sganga G, Shaller MD, Solomkin J (1998) 15 New Approaches in Fungal DNA Preparation from Whole Blood 335 Epidemiology, diagnosis and treatment of systemic Candida infection in surgical patients under intensive care. Intensive Care Med 24:206–216 Vincent JL, Sakr Y, Sprung CL, Ranieri VM, Reinhart K, Gerlach H, Moreno R, Carlet J, Le Gall JR, Payen D (2006) Sepsis in European intensive care units: results of the SOAP study. Crit Care Med 34:344–353 Wahyuningsih R, Freisleben HJ, Sonntag H-G, Schnitzler P (2000) Simple and rapid detection of Candida albicans DNA in serum by PCR for diagnosis of invasive candidiasis. J Clin Microbiol 38:3016–3021 Walsh TJ, Merz WG, Lee JW, Schaufele R, Sein T, Whitcomb PO, Ruddel M, Burns W, Wingard JR, Switchenko AC et al (1995) Diagnosis and therapeutic monitoring of invasive candidiasis by rapid enzymatic detection of serum D-arabinitol. Am J Med 99:164–172 Walsh TJ, Chanock SJ (1998) Diagnosis of invasive fungal infections: advances in nonculture systems. Curr Clin Top Infect Dis 18:101–153 Weinstein MP (2003) Blood culture contamination: persisting problems and partial progress. Clin Microbiol 41:2275–2278 Wellinghausen N, Wirths B, Essig A, Wassill L (2004) Evaluation of the Hyplex BloodScreen Multiplex PCR-Enzyme-linked immunosorbent assay system for direct identification of grampositive cocci and gram-negative bacilli from positive blood cultures. J Clin Microbiol 42:3147–3152 Wey SB, Mori M, Pfaller MA, Woolson RF, Wenzel RP (1988) Hospital-acquired candidaemia. The attributable mortality and excess length of stay. Arch Intern Med 148:2642–2645 White PL, Shetty A, Barnes RA (2003) Detection of seven Candida species using the Light-Cycler system. J Med Microbiol 52:229–238 White PL, Williams DW, Kuriyama T, Samad SA, Lewis MA, Barnes RA (2004) Detection of Candida in concentrated oral rinse cultures by real-time PCR. J Clin Microbiol 42:2101–2107 Wiesinger-Mayr H, Vierlinger K, Pichler R, Kriegner A, Hirschl AM, Presterl E, Bodrossy L, Noehammer C (2007) Identification of human pathogens isolated from blood using microarray hybridisation and signal pattern recognition. BMC Microbiol 7:78 Wilkinson CR, Bartlett R, Nurse P, Bird AP (1995) The fission yeast gene pmt1+ encodes a DNA methyltransferase homologue. Nucleic Acids Res 23:203–210 Wingard JR (2005) The changing face of invasive fungal infections in hematopoietic cell transplant recipients. Curr Opin Oncol 17:89–92 Wingard JR (1995) Importance of Candida species other than C. albicans as pathogens in oncology. Clin Infect Dis 20:115–125 Wong SF, Mak JW, Pook PC (2007) New mechanical disruption method for extraction of whole cell protein from Candida albicans. Southeast Asian J Trop Med Public Health 38:512–518 Yeo SF, Wong B (2002) Current status of nonculture methods for diagnosis of invasive fungal infections. Clin Microbiol Rev 15:465–484 Yong PV, Chong PP, Lau LY, Yeoh RS, Jamal F (2008) Molecular identification of Candida orthopsilosis isolated from blood culture. Mycopathologia 165:81–87 Chapter 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS and D1/D2 Regions of DNA Lidia Pérez-Pérez, Manuel Pereiro, and Jaime Toribio Abstract Yeasts of the genus Malassezia are known commensals of human beings and warm-blooded animals. Currently, they are considered emergent pathogen yeasts and have been described as causative agents of systemic opportunistic infections. An accurate identification of Malassezia spp. is of relevance to determine the role each species plays in the development of cutaneous and systemic infections. The taxonomy of Malassezia spp. has been a matter of discussion since the creation of the genus by Baillon in 1889. The recent development of molecular techniques has improved the classification of this genus, allowing a more accurate differentiation among different species. The taxonomic status is still under expansion, some new species have been appended recently and more will be probably added in the near future. However, descriptions of new species should be done in a standardized manner, including phenotypic and molecular features. We describe the current classification of Malassezia spp. yeasts based on the study and sequencing of ITS and D1/D2 gene of the rDNA and highlight the importance of these regions of the DNA as an easy tool for the identification of this genus. 16.1 Introduction Malassezia spp. is a common member of the skin flora which can become a pathogenic element after changes in the cutaneous microclimate. It has been implicated in the ethiopathogenic mechanisms of different skin diseases. Yeasts of the genus Malassezia have been classically identified by physiological tests L. Pérez-Pérez Department of Dermatology, University Hospital Complex of Vigo, C/Porriño 5, 36209 Vigo, Spain e-mail: lidiacomba@yahoo.es M. Pereiro and J. Toribio Department of Dermatology, Laboratory of Mycology, Faculty of Medicine, University Hospital Complex of Santiago de Compostela, C/ San Francisco S/N, 15706 Santiago de Compostela, Spain Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_16, # Springer-Verlag Berlin Heidelberg 2010 337 338 L. Pérez-Pérez et al. (Kaneko et al. 2007), which allow the identification of some but not all the Malassezia species recognized nowadays. The development of a range of molecular techniques based on the study of particular genes has been a definite step in the pathway of fungal identification (Gupta et al. 2004a; Canteros et al. 2007). They provide a more accurate identification, detecting slight differences and are also a very useful tool in the study of the taxonomy of the genus. They are also of great value in the understanding of the phylogenetic relationships and the analysis of the specific genetic variation of the species. The taxonomy and classification of the genus Malassezia are currently being updated and they will be probably enlarged with some other species in the near future. We present the classification of Malassezia spp. according to the sequencing of the D1/D2 domain and ITS regions of DNA and discuss the value of the novel molecular techniques as identification tools. 16.2 Description of the Article 16.2.1 Clinical Relevance of the Genus Malassezia Over the last few years, the implication of different yeasts in the development of many diseases in humans has become very relevant, particularly in some subgroups such as immunocompromised patients who may suffer from life-threatening widespread and severe infections. Malassezia spp. is a known commensal of warmblooded animals (including bears, monkeys, pigs, elephants, birds, horses, dogs, goats, sheep, cows, etc) and humans, mainly colonizing areas with a high density of sebaceous glands (scalp, face, ears, back, and chest). However, it can become a pathogenic agent under certain circumstances (Fig.16.1) such as high temperature and relative humidity, seborrhoeic constitution, hyperhidrosis, AIDS (Acquired Immunodeficiency Syndrome), hematological malignancy, organ transplants, intravascular devices, antitumoral therapy, corticoids, and broad spectrum antibiotics (Gupta et al. 2000, 2004b). The most common disease related to Malassezia spp. is pityriasis versicolor (Okuda et al. 1998; Crespo-Erchiga et al. 1999; Pereiro-Miguens 1999; Crespo et al. 2000a; Prohic and Ozegovic 2006; Krisanty et al. 2009) (Fig.16.2), which clinically manifests as multiple brownish, pinkish or whitish plaques with a mild desquamation that usually appear on areas rich in sebaceous glands. Some studies have suggested that M. globosa is the main causative agent of pityriasis versicolor (Crespo et al. 2000a); however, others have found M. furfur and M. sympodialis to be the most common species isolated in patients with this disease (Krisanty et al. 2009). Malassezia spp. may also play a role in the development of a number of skin diseases including pityriasis capitis, folliculitis, atopic dermatitis (Sugita et al. 2001, 2003b; Sandström et al. 2005) (Fig.16.3), seborrhoeic dermatitis (Gaitanis et al. 2006b; Prohic 2009), seborrhoeic blepharitis, confluent and reticulate papillomatosis (Gougerot and Carteaud syndrome) (Fig.16.4), transient acantholytic dermatosis, acne, psoriasis (Paulino et al. 2006; Ashbee 2006) (Fig.16.4), nodular 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 339 Fig. 16.1 Factors that can contribute to the pathogenic transformation of Malassezia spp Fig. 16.2 (a) Typical clinical appearance of pityriasis versicolor, with whitish macules on a male patient’s back. (b) A typical presentation of pityriasis versicolor, with atrophic patches and coexisting foliculitis on a female patient’s back hair infections, and onychomycosis (Gupta et al. 2004b). Midgley demonstrated that 72.5% of patients with seborrhoeic dermatitis showed precipitating antibodies to M. globosa (Midgley 2000). Crespo et al established that M. restricta and M globosa were the most common species isolated from 75 patients with seborrehoeic dermatitis (Crespo et al. 2000a). Folliculitis due to Malassezia spp. is a chronic process characterized by the presence of small pruritic follicular papules and pustules mainly distributed on the trunk and upper extremities. The role of Malassezia spp. in the development of atopic dermatitis is controversial. Some authorities have suggested that these yeasts might act as allergens mainly in those patients with lesions on their face and neck (Faergemann 1999). Confluent and reticulated 340 L. Pérez-Pérez et al. Fig. 16.3 Atopic eczema on a male patient’s forearm Fig. 16.4 (a) Clinical appearance of reticulate papillomatosis of Gougerot and Carteaud on a male patient’s back. (b) Typical plaques of psoriasis on a male patient’s back papillomatosis is an uncommon disorder clinically characterized by greyish to brownish papules on the trunk or abdomen. The relevance of the identification of Malassezia yeasts in patients with this condition and also in patients with psoriasis is yet unknown. Malassezia spp. has also been suggested to play a role in fungal nail disease (Midgley 2000) and has also been identified as the causative agent for different severe extracutaneous diseases (Gueho et al. 1987), particularly in patients with underlying diseases or predisposing factors, such as pneumonia, mastitis, sinusitis, malignant otitis externa, abscesses, meningitis (Ashbee 2007), catheterrelated fungaemia (Juncosa Morros et al. 2002), and peritonitis (Aspı́roz et al. 1997). M. furfur and to a lesser extent M. pachydermatis have been implicated in the development of severe systemic disease in neonates. Malassezia spp. can also cause many skin diseases in animals such as otitis externa (Crespo et al. 2000b; Hirai et al. 2004), dermatitis (Cabañes et al. 2005), alopecia, ulcerous lesions, pruritus, and liquenification (Chen and Hill 2005). 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 341 16.2.2 The Diagnosis of the Genus Malassezia The identification of the genus Malassezia and the differentiation among species have been classically made on the basis of phenotypical and physiological features (Table 16.1) such as the morphology of the colonies, size and shape of the cells (Fig.16.5), and nutritional requirements (Guillot et al. 1996; Kaneko et al. 2007). The macroscopic appearance of the colonies varies among the species. The colonies are generally small, creamy to yellowish in color, with a smooth or rough surface (Fig.16.6) and are composed of rounded, cylindrical or oval cells (Gueho et al. 1996). A characteristic physiological feature of these yeasts is their lipid-dependency, which may vary among the species. The lipid-dependent species need long chain fatty acids to grow; meanwhile those non lipid-dependent species may grow in common culture media containing short chain fatty acids. Except for M. pachydermatis, yeasts of the genus Malassezia are all lipid-dependent. However, different authors have described isolates of M. pachydermatis, phenotypically identified as lipid-dependent, this fact still being a matter of controversy. (Cafarchia et al. 2007). Some authorities have suggested it would be the result of a process of differentiation or adaptation to a particular host. The lipid-dependent species are usually isolated from human skin whereas M. pachydermatis has been isolated mainly from birds and mammals. The former have been associated with several diseases and the latter causes chronic dermatitis and otitis externa in animals and also nosocomial infection in humans (Cafarchia et al. 2007). Other physiological features comprise the catalase reaction, assimilation of different polyethilenesorbitane esters (Tween 20, 40, 60, 80), enzymatic activity (esterase, esterase–lipase, N-phosphohydrolase, acid phosphatase, alkaline phosphatase, phospholipase (Cafarchia et al. 2008)), and tolerance to temperature. Guillot et al developed a physiologic algorithm for the identification of Malassezia species based on lipid assimilation and other phenotypic features (Guillot et al. 1996). Kaneko et al. developed a culture-based system for the identification of Malassezia species which allowed an easy identification of nine different species (particularly M. furfur, M. globosa and M. restricta) with a rate of concordance with molecular analysis of 98.1% (Kaneko et al. 2007). The atypical assimilation of Tween 80 has been recently found to be of interest for the identification of M. furfur (González et al. 2009). These physiological tests, however, present some limitations and difficulties: they are time-consuming, their results are variable, and sometimes display an inadequate taxonomic value. The classical methods do not allow a certain identification of all the species and thus are not enough for classification purposes. 16.2.3 Current State of the Classification of Malassezia spp. Until 1990, only three species of Malassezia were recognized, M. furfur (Robin) Baillon, M. pachydermatis (Weidman) Dodge, and M. sympodialis 342 L. Pérez-Pérez et al. Table 16.1 Physiological characteristics of the main Malassezia species: 1 ¼ M. furfur; 2 ¼ M. pachydermatis; 3 ¼ M. sympodialis; 4 ¼ M. globosa; 5 ¼ M. obtusa; 6 ¼ M. restricta; 7 ¼ M. slooffiae; 8 ¼ M. dermatis; 9 ¼ M. nana; 10 ¼ M. japonica; 11 ¼ M. yamatoensis. (Guillot et al. 1996; Hirai et al. 2004; Sugita et al. 2003b) Features 1 2 3 4 5 6 7 8 9 10 11 Culture in Sabouraud medium at 32 C + (no lipid supplementation) Culture in Dixon medium at 40 C + + + + + +/ Catalase reaction + +/ + + + + + + + + +/ Lipid assimilation: +/ + + + + + +/ +/ – Tween 20 (10%) + + + +/ + + – Tween 40 or 60 (0.5%) + + + NTb – Tween 80 (0.1%) + Va – Cremophor EL Esculine hydrolisis V + + NT NT NT + + NT + + NT NT Precipitate in Dixon medium a V-variable b NT-not tested 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 343 Fig. 16.5 Shape of Malassezia globosa cells Fig. 16.6 Small, whitish and creamy colonies of Malassezia pachydermatis (Simmons and Guého) (Gupta et al. 2000). In 1996, Guého et al updated the genus Malassezia according to morphologic, ultrastructural, physiologic, and molecular criteria and established a new classification comprising seven well-defined species (Gueho et al. 1996). Recently, on the basis of genetic studies of a number of different isolates from humans and animals, four new species have been suggested or proposed: M. dermatis (Sugita et al. 2002), M. nana (Hirai et al. 2004), M. japonica (Sugita et al. 2003b), and M. yamatoensis (Sugita et al. 2004) (Table 16.2). Nell et al. isolated a new species from horses, M. equi, which has not been formally recognized (Nell et al. 2002), as its description was not provided and there is not any type specimen currently available. Cabañes et al recently described two new lipid-dependent species named Malassezia caprae sp. nov. and Malassezia equina sp. nov., isolated from healthy goats and horses, respectively (Cabañes et al 2007). These species seem to be closely related to M. sympodialis, M. Dermatis, and M. nana. M. equina, M. caprae, M. Nana, and M. pachydermatis are associated with animals; the remaining species are part of the normal human flora and are also associated with human pathologies (González et al. 2009). Several molecular techniques have been introduced recently in the field of mycology for the study and identification of fungi. Those applied to study the genus Malassezia include RADP (Random amplification of polymorphic DNA) 344 Table 16.2 Currently accepted and proposed Malassezia species L. Pérez-Pérez et al. Malassezia species M furfur M. pachydermatis M. sympodialis M globosa M. slooffiae M. obtusa M. restricta M. dermatis M. japonica M. nana M. yamatoensis M. caprae sp nov M. equina sp nov Author, year (Robin) Baillon (1889) (Weidman) Dodge (1935) Simmons and Guého (1990) Midgley et al. (1996) Guillot et al. (1996) Gueho et al. (1996) Gueho et al. (1996) Sugita et al. (2002) Sugita et al. (2003) Hirai et al. (2004) Sugita et al. (2004) Cabañes et al. (2007) Cabañes et al. (2007) (Castellá et al. 2005; Hossain et al. 2007), PFGE (pulsed-field gel electrophoresis) (Boekhout et al. 1998), sequencing of the chitin synthase 2 gene (Kano et al. 1999), sequencing of the large subunit of mitochondrial rRNA (Yamada et al. 2003), AFLP (Amplified Length polymorphisms) (Gupta et al. 2004a), PCR of the LSU of the rRNA and digestion with restriction enzymes (Guillot et al. 2000), multiplex-real time PCR (Paulino et al. 2008), PCR-REA (Polymerase Chain Reaction and Restriction Enzyme Analysis) (Giusiano et al. 2003), DGGE (denaturing gradient gel electrophoresis) (Theelen et al. 2001), RFLP (restriction fragment polymorphisms) (Mirhendi et al. 2005), sequencing of the LSU of the rDNA or the ITS region (Sugita et al. 2003a; Cafarchia et al. 2007). The most recent method developed for the identification of Malassezia spp. is a bead suspension array that uses species and group specific probes analyzed by flow cytometry (Dı́az et al. 2006). Cafarchia et al suggested the use of multilocus sequencing for the identification of and differentiation among species or genotypes which are phenotypically difficult to characterize (Cafarchia et al. 2007). Gaitanis et al published recently an interesting and detailed review on the methodology for Malassezia typing, distinguishing two main groups of techniques: those focused on targeted PCR amplifications of selected sequences and subsequent search of mutations and random PCR-amplification of polymorphic DNA or redundant sequences within the genome (Gaitanis et al. 2009). Most of these methods are expensive and time consuming. In contrast, PCR techniques are easy to perform and provide a rapid identification of Malassezia species (Affes et al. 2008). Sequencing of the D1/D2 domain and the ITS regions (Gaitanis et al. 2006a; Makimura et al. 2000) of the DNA is a useful and accurate procedure for the identification and classification of different fungi (Abliz et al. 2004; León-Mateos et al. 2006). The rDNA genes of Malassezia spp. are constituted by the 5S, 5.8S, 18S (small), and 26S (large) subunits (Fig.16.7). There are other two regions inserted between the subunits: the ITS region (Internal Transcriber Spacer) and the IGS region (Intergenic Spacer). Both of them are divided into two subregions. The D1/D2 region of the DNA which encodes for the ribosomal large subunit (LSUrDNA) is located in the 50 end of the large subunit (26S) of the ribosomal 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 345 Fig. 16.7 Structure of the rDNA gene of Malassezia spp DNA (Sugita et al. 2003a). These regions are highly conserved in a particular species and vary among different species. rDNA is a multicopy gene that includes several regions not encoding for proteins: ITS 1, ITS 2, and IGS, which are highly conserved regions. The D1/D2 region of the DNA encoding for the ribosomal large subunit (LSUrDNA) is currently considered very useful for the identification of the vast majority of fungi of medical relevance (Kurtzman and Robnett 1997; Fell et al. 2000). Most of the species can be identified by analysis of the D1/D2 domain. In fact, no other genetic region with a higher ability to discriminate species has been described to date (Abliz et al 2004). However, the study of closely related species or strains requires sequencing of the ITS region (Cabañes et al. 2005). The sequencing of the ITS regions and D1/D2 domain and subsequent studies of the sequences obtained showed that the phylogenetic trees constructed with the ITS region sequencing matched those obtained with D1/D2 domain sequencing (Gupta et al. 2004a). This method allows an accurate identification and differentiation of the current Malassezia species, which appear in the trees grouped in different clusters. Small differences among strains from the same species can be found sometimes and thus a particular species can be considered in terms of a “species complex.” Our group (Laboratory of Mycology, Department of Dermatology, Faculty of Medicine, Santiago de Compostela) conducted studies on the D1/D2 and ITS regions of isolates from different species of Malassezia, in an attempt to analyze their phylogenetic relations (unpublished data). We studied a total of 28 strains from six different species of the genus Malassezia (M. pachydermatis, M. sympodialis, M. furfur, M. restricta, M. globosa, M. sloffiae) (Table 16.3). These isolates were collected from dogs, pigs, healthy subjects, and patients with pytiriasis versicolor and were part of a 63-strain collection previously used by our group to carry out a study on the LSU and ITS regions (unpublished data). The strains were cultured in Saboraud and modified Dixon’s media supplemented with cyclohemixida and chloramfenicol following standard procedures and were first identified on the basis of their phenotypic and physiological features. DNA extraction was subsequently performed as described by Hillis (Hillis et al. 1996), after digestion of the cellular wall. The ITS region and D1/D2 domain were amplified by PCR using respectively the fungal oligonucleotides (ITS-5, ITS-2, ITS-3, ITS-4) and (NL-1, NL-2, NL-3, NL-4) (Table 16.4) synthesized by Sigma 346 Table 16.3 Details of the strains sequenced in our study L. Pérez-Pérez et al. Origin Animal, pig Animal, pig Animal, dog Animal, dog Animal, dog Animal, dog Animal, dog Animal, dog Animal, dog Animal, dog Animal, dog Animal, dog Animal, pig Animal, pig Animal, pig Animal, pig Animal, pig Animal, pig Human Human Human Animal, pig Animal, pig Human Human Human Human Animal, pig Strain M. furfur 26 M. furfur 28 M. pachydermatis 4 M. pachydermatis 6 M. pachydermatis 10 M. pachydermatis 34LD M. pachydermatis 21 M. pachydermatis 75LD M. pachydermatis 13LD M. pachydermatis 11 M. pachydermatis 75 M. pachydermatis 107LD M. sympodialis 12A M. sympodialis 13A M. sympodialis 20 M. sympodialis 21 M. sympodialis 22A M. sympodialis 25A M. sympodialis 01022043 M. sympodialis 043943B M. sympodialis 039371A M. sympodialis 17A M. sympodialis 32A M. sympodialis CBS 7222 M. globosa 01034998 M. globosa 010425748 M. restricta VCA 585 M. slooffiae 3A Code MFU01 MFU02 MPA03 MPA04 MPA05 MPA06 MPA07 MPA08 MPA09 MPA10 MPA11 MPA12 MSY13 MSY14 MSY15 MSY16 MSY17 MSY18 MSY19 MSY20 MSY21 MSY22 MSY23 MSY24 MGL25 MGL26 MRE27 MSL28 Table 16.4 Primers used for the amplification of the D1/D2 domain and the ITS regions of Malassezia spp Tm ( C) Region Primer Sequence (50 –30 ) D1/D2 NL-1 gca tat caa taa gcg gag gaa aag 650 33 NL-4 m ggt ccg tgt ttc aag acg 610 79 ITS ITS1 tcc gta ggt gaa ccg cgc 65 ITS5 gga agt aaa gtc gta aca agg 63 ITS2 gct gcg ttc ttc atc gat gc 62 ITS3 gca tcg atg aag aac gca gc 62 ITS4 tcc tcc gct tat tga tat gc 58 Genosys labs (Sigma Genosys Ltd, Sigma-Aldrich House, Haverhill, UK). These primers had already been successfully used in previous studies of these regions (Hirai et al. 2002; Sugita et al. 2002, 2003a; Martin and Rygiewicz 2005; Cabañes et al. 2005). We extracted a number of sequences of different Malassezia species from the Genbank database (Table 16.5) with comparative purposes and to study their phylogeny. 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 347 Table 16.5 Sequences extracted from the GenBank database and their accession numbers Species Strain D1/D2 domain ITS region accession number accesion number M. sympodialis WBC2 AY 387254 – LMB3 AY 387268 – KEB1 AY 387271 – JF05 AY 387276 – 98F AY 387291 – MA231 AY 743609 AY743648 MA80.CBS9967 AY 743618 AY743647 MA125 AY 743619 AY743646 MA73.CBS9968 AY 743627 AY743640 MA 477 AY 743628 AY743639 MA88CBS9986 – AY743645 MA419 – AY743653 M. pachydermatis GENOTYPE A DQ915500 – GENOTYPE B DQ915501 – GENOTYPE C1 DQ915502 – CBS1879 AY387235 – CBS1919 AY387236 – CBS1885 AY387237 – CBS1884 AY387238 – CBS1879 AY743605 AY743637 AFTOL-ID856 AY745724 – CBS1879 AJ249952 – CBS1879 – AB118941 IFM52772 – AB118939 IFM52748 – AB118940 IFM52755 – AB118937 M. furfur CBS1878 AY743602 AY743634 CBS7019 AY743603 AY743635 CBS1878 AY387196 – M. yamatoensis M9985 AB12563 AB125261 M9986 AB12564 M. japonica CBS9431 EF140672 M9976 AB105199 AB105199 M. slooffiae CBS7956 AY743606 AY743633 AJ249956 – TV1 AY387249 – M. globosa CBS7966 AJ249951 – AY387228 – AY743604 AY743630 M. restricta CBS7877 AJ249950 – AY387239 – AY743607 AY743636 M. obtusa CBS7876 AY743629 AY743631 CBS7968 AY387234 – M. dermatis M9927 AB070361 AB070356 M9930 AB070364 – 348 L. Pérez-Pérez et al. The sequences obtained were analyzed with the software DNAsis, CLC Free Workbench v. 4.0.2 and MEGA version 3.1 (Kumar et al 2004).The clustal alignment of our sequences and those extracted from the GenBank database was performed with the software CLC Free Workbench v.4.0.2. Subsequent comparative studies of the sequences obtained were conducted to analyze inter and intraspecies dissimilarities. The MEGA package, version 3.1, was used to perform a maximum parsimony analysis with 1,000 bootstrap replicates. The molecular phylogenetic trees of the ITS and D1/D2 regions of the 26s rRNA gene sequences were constructed with MEGA v3.1 using the maximum parsimony method. The D1/D2 domain tree in Fig. 16.8 shows four well-defined phylogenetic clusters: cluster I comprises M. sympodialis strains, which seem to be closely related to M. dermatis strains and display intraspecies diversity. Strains of M. sympodialis from animals and humans tend to appear grouped in the tree. Cluster II includes M. obtusa, M. yamatoensis, M. Japonica, and M. furfur. M. yamatoensis and M. japonica are both from human origin and appear classified within the same subcluster. All the M. pachydermatis strains were isolated from animals (dogs) and are grouped in cluster III. The vast majority of our strains are grouped together, closely related to those from the GenBank database. Cluster IV includes M. globosa, M. Restricta, and M. slooffiae, which are clearly separated into three independent subgroups. The phylogenetic tree constructed with the sequences of the ITS region of our strains and those selected from the GenBank database is shown in Fig. 16.9. As it occurs with the tree obtained with the D1/D2 domain sequences, four main clusters are identified: cluster I includes M. sympodialis, M. Dermatis, and M. yamatoensis; cluster II comprises all the strains of M. pachydermatis, which seem to share a common root with M. sympodialis. All of them appear grouped together. Cluster III includes strains of M. japonica, M. Obtusa, and M. furfur. M. slooffiae, M. Globosa, and M. restricta constitute cluster IV. The phylogenetic relationships displayed in the trees constructed with the sequences of ITS and D1/D2 regions are very similar. 16.3 Conclusions and Future Line of Investigation The molecular methods have provided strong evidence to support the current classification of Malassezia spp. and have contributed much to improve the understanding and knowledge of the genus (Gaitanis et al. 2009). However, they present some limitations. Firstly, they need specific equipment and trained staff and are expensive in comparison to the classical methods of identification. Moreover, on clinical grounds, for most routine clinical mycology laboratories there is little need to speciate the isolates recovered from skin samples. Although it is of epidemiological interest to determine the species of Malassezia associated with particular 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS Fig. 16.8 D1/D2 phylogenetic tree (See text). The numbers at branch points represent the percentages of 1,000 bootstrapped datasets that supported the specific internal branches 349 Msy AY743627 Msy AY387276 Mder AB070361 99 Mder AB070364 MSY24 Msy AY387268 MSY20 Msy AY387254 2 MSY16 Msy AY743628 37 MSY15 MSY21 Cluster I Msy AY387271 Msy AY387291 96 Msy AY743619 34 27 Msy AY743618 Msy AY743609 MSY14 35 24 MSY17 MSY22 MSY13 99 MSY19 MSY23 MSY18 43 99 Mob AY387234 Mob AY743629 99 Myam AB12563 Myam AB12564 38 96 Mjap EF140672 Cluster II 99 Mjap AB105199 MFU02 91 MFU01 Mfu AY743602 86 Mfu AY743603 Mfu AY387196 68 MPA05 60 MPA03 MPA10 Mpa DQ915502 MPA09 Mpa DQ915501 MPA12 Mpa AY387237 MPA08 98 MPA04 Cluster III MPA07 MPA11 MPA06 Mpa AY387238 Mpa AY743605 46 Mpa AJ249952 Mpa DQ915500 84 Mpa AY387235 Mpa AY745724 Mpa AY387236 Mgl AY387228 MGL26 99 Mgl AJ249951 MGL25 Mgl AY743604 Mre AY743607 40 Cluster IV 99 MRE27 Mre AJ249950 Mre AY387239 Msl AY743606 31 Msl AY387249 99 Msl AJ249956 MSL28 5 350 L. Pérez-Pérez et al. Fig. 16.9 ITS phylogenetic tree (See text). The numbers at branch points represent the percentages of 1,000 bootstrapped datasets that supported the specific internal branches 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 351 diseases, this is beyond the scope of most laboratories in a routine clinical setting and thus, most of these methods are used in laboratories for only investigative purposes. Because of the time it takes to culture Malassezia and the realization that no single medium can reliably recover all the species, several groups have developed methods for the molecular analysis of fungus directly from the skin without prior culture (Ashbee 2007) or from a range of commercial media used in microbiology laboratories (Pryce et al. 2006). This would be an interesting alternative, as it shortens the identification process and makes it easier. Anyway, the results obtained by molecular methods should be carefully interpreted under the light of the classical phenotypic methods of identification and clinical findings, as the molecular classification is based on genetic sequences that encode proteins and these are not directly related to the phenotypic characteristics of the species. The current available analyzes study highly conserved regions of the nuclear DNA encoding for the ribosomal subunits. Differences among the sequences obtained seem to overlap with clear phenotypic differences observed among the species. The dissimilarities found in these regions may affect the intergenic spaces, which do not encode for protein (intron) or the spaces encoding for a particular subunit. Intergenic spaces are of high value, as they include transcription regulating regions. Along the evolution of the species, these spaces tend to shorten due to deletions in the sequences and thus it is relatively common to find differences. However, it can be difficult to establish whether these differences merit to be considered a different species or are simply variations of a particular species. The regions encoding for functional proteins (exon) tend to remain conserved and therefore significant differences in them are more difficult to observe. It would also be of interest to analyze in future studies whether the differences among these two regions (intron and exon) are concordant. We tried to compare sequences of the ITS region and the D1/D2 domain of the same strain, as we thought the comparison between these two regions of the DNA would be stronger. It was not possible to do so with all the strains, as many of them do not have sequences available for both regions in the GenBank database. We consider it would be of interest to enlarge in the future the databases with strains which were sequenced in both the ITS and D1/D2 domain, in order to compare this two regions properly and analyze their relationships. ITS region and D1/D2 domain are much conserved. Small differences between them are of high relevance, particularly in the D1/D2 domain, as it is more conserved than the ITS region. We consider of interest the classification of the different species on the basis of the sequences of both regions. Some species showed high diversity in the phylogenetic trees, thus suggesting that the existence of close and different species not yet identified might be a strong possibility. We observed that M pachydermatis and M sympodialis particularly showed higher genetic diversity in the regions we studied and we suggest it would be interesting to run a larger study on these species, as some of the strains currently included may be recognized as new species in the future. A deeper knowledge of all the Malassezia species and their biology is not only necessary to fully understand the mechanism of many skin diseases, but also to 352 L. Pérez-Pérez et al. improve and complement the different treatment options currently available for these disorders. Uploading new sequences of different species of Malassezia in the GenBank database is of great interest in enlarging the number of sequences available for studies in the near future. Larger investigations correlating the molecular dissimilarities among species and their phenotypical and biological characteristics are the key to fully understand the real meaning of the findings in the molecular methods and their impact in the daily routine. Acknowledgments We are very grateful to l l l l Dr. Carmen Paredes Suárez (Dermatology Unit; Hospital Virxe da Xunqueira, Cee, Spain), who kindly provided her personal data on the restriction enzyme analysis of the ITS region. Dr Álvaro León Mateos (Department of Dermatology; Hospital POVISA, Vigo, Spain) for his advise on the writing of this manuscript. Dr. Margarita Garau and Dr Amalia del Palacio (Department of Microbiology; Hospital 12 de Octubre, Madrid, Spain), for providing some of the strains for our study. Dr. Gillian Midgley (Department of Medical Mycology, St John’s Institute of Dermatology, St Thomas’ Hospital, London, UK) for her disinterested contribution to the understanding of the genus Malassezia. This study was supported by the grant PGIDT01PX120810PR from the Local Government of Galicia, Spain. References Abliz P, Fukushima K, Takizawa K, Nishimura K (2004) Identification of pathogenic dematiaceous fungi and related taxa based on large subunit ribosomal DNA D1/D2 domain sequence analysis. FEMS Immunol Med Microbiol 40:41–49 Affes M, Ben Salah S, Makni F, Sellami H, Ayadi A (2008) Molecular identification of Malassezia species isolated from dermatitis affections. Mycoses 52:251–256 Ashbee HR (2006) Recent developments in the immunology and biology of Malassezia species. FEMS Immunol Med Microbiol 47:14–23 Ashbee HR (2007) Update on the genus Malassezia. Med Mycol 45:287–303 Aspı́roz MC, Moreno LA, Rubio MC (1997) Taxonomı́a de Malassezia furfur: estado de la cuestión. Rev Iberoam Micol 14:147–149 Boekhout T, Kamp M, Guého E (1998) Molecular typing of Malassezia species with PFGE and RAPD. Med Mycol 36:365–372 Cabañes FJ, Hernández JJ, Castellá G (2005) Molecular analysis of Malassezia sympodialisrelated strains from domestic animals. J Clin Microbiol 43:277–283 Cabañes FJ, Theelen B, Castellá G, Boekhout T (2007) Two new lipid-dependent Malassezia species from domestic animals. FEMS Yeast Res 7:1064–1076 Cafarchia C, Latrofa M, Testini G et al (2007) Molecular characterization of Malassezia isolates from dogs using three distinct genetic markers in nuclear DNA. Mol Cell Probes 21:229–238 Cafarchia C, Gasser R, Latrofa M, Parisi A, Campbell B, Otranto D (2008) Genetic variants of Malassezia pachydermatis from canine skin. Body distribution and phospholipase activity. FEMS Yeasts Res 8:451–459 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 353 Canteros CE, Rivas MC, Lee W, Perrotta D, Bosco-Borgeat ME, Davel G (2007) Concordancia entre caracterı́sticas fenotı́picas y PCR-REA en la identificación de especies de Malassezia. Rev Iberoam Micol 24:274–278 Castellá G, Hernández J, Cabañes J (2005) Genetic typing of Malassezia pachydermatis from different domestic animals. Vet Microbiol 108:291–296 Chen T, Hill PB (2005) The biology of Malassezia organisms and their ability to induce immune responses and skin disease. Vet Dermatol 16:4–26 Crespo-Erchiga V, Ojeda-Martos A, Vera-Casano A, Crespo-Erchiga A, Sánchez-Fajardo F (1999) Aislamiento e identificación de Malassezia spp. en pitiriasis versicolor, dermatitis seborreica y piel sana. Rev Iberoam Micol 16:S16–S21 Crespo V, Ojeda A, Vera A, Crespo A, Sánchez F (2000a) Malassezia globosa as the causative agent of pityriasis versicolor. Br J Dermatol 143:799–803 Crespo M, Abarca M, Cabañes FJ (2000b) Otitis externa associated with Malassezia sympodialis in two cats. J Clin Microbiol 38:1263–1266 Dı́az MR, Boekhout T, Theelen B et al (2006) Microcoding and flow cytometry as a high-throughput fungal identification system for Malassezia species. J Med Microbiol 55:1197–1209 Faergemann J (1999) Pityrosporum species as a cause of allergy and infection. Allergy 54:93–99 Fell JW, Boekhout T, Fonseca A, Scorzetti G, Scorzetti G, Statzell-Tallman A (2000) Biodiversity and systematics of Basidiomycetous yeasts as determined by large-subunit rDNA D1/D2 domain sequence analysis. Int J Syst Evol Microbiol 50:1351–1371 Gaitanis G, Robert V, Velegraki A (2006a) Verifiable single nucleotide polymorphisms of the internal transcribed spacer 2 region for the identification of 11 Malassezia species. J Dermatol Sci 43:214–217 Gaitanis G, Velegraki A, Alexopouos EC, Chasapi V, Tsigonia A, Katsambas A (2006b) Distribution of Malassezia species in pityriasis versicolor and seborrhoeic dermatitis in Greece. Typing of the major pityriasis versicolor isolate M. globosa. Br J Dermatol 154:854–859 Gaitanis G, Bassukas I, Velegraki A (2009) The range of molecular methods for typing Malassezia. Curr Opin Infect Dis 22:119–125 Giusiano G, Bustillo S, Mangiaterra M, Deluca G (2003) Identificación de especies de Malassezia por PCR-REA. Rev Argent Microbiol 35:162–166 Gueho E, Simmons R, Pruitt W, Meyer S, Ahearn D (1987) Association of Malassezia pachydermatis with systemic infections of humans. J Clin Microbiol 25:1789–1790 Gueho E, Midgley G, Guillot J (1996) The genus Malassezia with description of four new especies. Antonie van Leeuwenhoek 69:337–355 Guillot J, Gueho E, Lesourd M, Midgley G, Chevrier G, Dupont B (1996) Identification of Malassezia species. J Mycol Med 6:103–110 Guillot J, Deville M, Berthelemy M, Provost F, Guého E (2000) a single PCR-restriction endonuclease analysis for rapid identification of Malassezia species. Lett Appl Microbiol 31:400–403 González A, Sierra R, Cárdenas ME, Grajales A, Restrepo S, Cepero de Garcı́a MC, Celis A (2009) Physiological and molecular characterization of atypical isolates of Malassezia furfur. J Clin Microbiol 47:48–53 Gupta AK, Kohli Y, Summerbell RC (2000) Molecular differentiation of seven Malassezia species. J Clin Microbiol 38:1869–1875 Gupta AK, Boekhout T, Theelen B, Summerbell R, Batra R (2004a) Identification and typing of Malassezia species by amplified length polymorphism and sequence analysis of the internal transcribed spacer and large-subunit regions of ribosomal DNA. J Clin Microbiol 42:4253–4260 Gupta AK, Batra R, Bluhm R, Boekhout T, Dawson TL Jr (2004b) Skin diseases associated with Malassezia species. J Am Acad 51:785–798 Hillis DM, Mable BK, Larson A, Davis SK, Zimmer EA (1996) Nucleic acids IV: sequencing and cloning. In: Hillis DM, Moritz G, Mable BK (eds) Molecular systematics, 2nd edn. Sinauer, Canada, pp 343–356 354 L. Pérez-Pérez et al. Hirai A, Kano R, Makimura K, Yasuda K, Konishi K, Yamagushi H (2002) A unique isolate of Malassezia from a cat. J Vet Med Sci 64:957–959 Hirai A, Kano R, Makimura K, Robson E, Soares J, Lachance MA et al (2004) Malassezia nana sp. nov., a novel lipid-dependent yeast species isolated from animals. Int J Syst Evol Microbiol 54:623–627 Hossain H, Landgraff V, Weiss R et al (2007) Genetic and biochemical characterization of Malassezia pachydermatis with particular attention to pigment-producing subgroups. Med Mycol 45:41–49 Juncosa Morros T, González-Cuevas A, Alayeto-Ortega J, Muñoz-almagro C, Moreno-Hernando J, Gené-Giralt A, Latorre-Otı́n C (2002) Colonización cutánea neonatal por Malassezia spp. An Esp Pediatr 57:452–456 Kaneko T, Makimura K, Abe M et al (2007) Revised culture-based system for identification of Malassezia species. J Clin Microbiol 45:3737–3742 Kano R, Aizawa T, Nakamura Y, Watanabe S, Hasegawa A (1999) Chitin synthase 2 gene sequence of Malassezia species. Microbiol Immunol 43:813–815 Krisanty RI, Bramono K, Made Wisnu I (2009) Identification of Malassezia species from pityriasis versicolor in Indonesia and its relationship with clinical characteristics. Mycoses 52:257–262 Kumar S, Tamura K, Nei M (2004) MEGA3: integrated software for molecular evolutionary genetics analysis and sequence alignment. Brief Bioinform 5:150–163 Kurtzman CP, Robnett CJ (1997) Identification of clinically important Ascomycetous yeasts based on nucleotide divergence in the 50 end of the large –subunit (26S) ribosomal DNa gene. J Clin Microbiol 35:1216–1223 León-Mateos A, Paredes Suárez C, Pereiro M, Toribio J (2006) Study of the ITS region in an atypical isolate and comparison with six species of Microsporum. Mycoses 49:452–456 Makimura K, Tamura Y, Kudo M, Uchida K, Saito H, Yamaguchi H (2000) Species identification and strain typing of Malassezia species stock strains and clinical isolates based on the DNA sequences of nuclear ribosomal internal transcribed spacer 1 regions. J Med Microbiol 49:29–35 Martin K, Rygiewicz PT (2005) Fungal-specific PCR primers developed for analysis of the ITS region of environmental DNA extracts. BMC Microbiol 5:28–38 Midgley G (2000) The lipophilic yeasts: state of the art and prospects. Med Mycol 38:9–16 Mirhendi H, Makimura K, Zomorodian K, Yamada T, Sugita T, Yamaguchi H (2005) A simple PCR-RFLP method for identification of 11 Malassezia species. J Microbiol Methods 61:281–284 Nell A, James SA, Bond CJ et al (2002) Identification and distribution of a novel Malassezia species yeast on normal equine skin. Vet Rec 150:395–398 Okuda C, Ito M, Naka W et al (1998) Pityriasis versicolor with a unique clinical apperance. Med Mycol 36:331–334 Paulino L, Tseng C, Strober B, Blaser M (2006) Molecular analysis of fungal microbiota in samples from healthy human skin and psoriatic lesions. J Clin Microbiol 44:2933–2941 Paulino L, Tseng C, Blaser M (2008) Analysis of Malassezia microbiota in healthy sperficial human skin and in psoriatic lesions by multiplex real-time PCR. FEMS Yeast Res 8:460–471 Pereiro-Miguens M (1999) Situación actual de las infecciones por Malassezia. Piel 14:76–87 Prohic A, Ozegovic L (2006) Malassezia species isolated from lesional and non-lesional skin in patients with pityriasis versicolor. Mycoses 50:58–63 Prohic A (2009). Distribution of Malassezia species in seborrheic dermatitis: correlation with patients’ cellular immune status. Mycoses (In press). doi:10.1111/j.1439-0507.2009.01713.x Pryce T, Palladino S, Price D et al (2006) Rapid identification of fungal pathogens in BacT/ ALERT, BACTEC and BBL MGIT media using polymerase chain reaction and DNA sequencing of the internal transcribed spacer regions. Diagn Microbiol Infect Dis 54:289–297 Sandström MH, Tengvall M, Johansson C, Bartosik J, Bäck O, Särnhult T et al (2005) The prevalence of Malassezia yeasts in patients with atopic dermatitis, seborrhoeic dermatitis and healthy controls. Acta Derm Venereol 85:17–23 16 Classification of Yeasts of the Genus Malassezia by Sequencing of the ITS 355 Sugita T, Suto H, Unno T et al (2001) Molecular analysis of Malassezia microflora on the skin of atopic dermatitis patients and healthy subjects. J Clin Microbiol 39:3486–3490 Sugita T, Takashima M, Shinoda T, Suto H, Unno T, Tsuboi R, Ogawa H, Nishikawa A (2002) New yeast species, Malassezia dermatis, isolated from patients with atopic dermatitis. J Clin Microbiol 40:1363–1367 Sugita T, Kodama M, Saito M et al (2003a) Sequence diversity of the intergenic spacer region of the rRNA gene of Malassezia globosa colonizing the skin of patients with atopic dermatitis and healthy individuals. J Clin Microbiol 41:3022–3027 Sugita T, Takashima M, Kodama M, Tsuboi R, Nishikawa A (2003b) Description of a new yeast species, Malassezia japonica, and its detection in patients with atopic dermatitis and healthy subjects. J Clin Microbiol 41:4695–4699 Sugita T, Tajima M, Takashima M et al (2004) A new yeast, Malassezia yamatoensis, isolated from a patient with seborrheic dermatitis, and its distribution in patients and healthy subjects. Microbiol Immunol 48:579–583 Theelen B, Silvestri M, Guého E, Van Belkum A, Boekhout T (2001) Identification and typing of Malassezia yeasts using amplified fragment length polymorphism (AFLP), random amplified polymorphic DNA (RAPD) and denaturing gradient gel electrophoresis (DGGE). FEMS Yeast Res 1:79–86 Yamada Y, Makimura K, Ueda K et al (2003) DNA base alignment and taxonomic study of the genus Malassezia based upon partial sequences of mitochondrial large subunit ribosomal RNA gene. Microbiol Immunol 47:475–478 Chapter 17 DNA-Based Detection of Human Pathogenic Fungi: Dermatophytes, Opportunists, and Causative Agents of Deep Mycoses Lorenza Putignani, Silvia D’Arezzo, Maria Grazia Paglia, and Paolo Visca Abstract The affordability of modern molecular biology tools and the availability of whole genome sequences have brought substantial improvement in research on pathogenic fungi and diagnosis of fungal infection. Molecular methods have resolved many critical aspects of mycological diagnosis by (1) providing specieslevel identification of fungi through sequencing of suitable taxonomic markers; (2) shortening of the time required for microbiological confirmation of life-threatening fungal infections; and (3) tracing the molecular epidemiology of fungal diseases. Nucleic acids-based methods are less subjective than microscopy- or culture-based methods and unaffected by fungal growth conditions, thus capable of discriminating between phenotypically undistinguishable species. This chapter focuses on the contribution of DNA-based techniques to the identification of clinically important fungi such as Aspergillus, Blastomyces, Candida, Coccidioides, Cryptococcus, Dematiaceous fungi, Fusarium, Histoplasma, Trichosporon, Zygomycetes, and Dermatophytes. Because of their excellent performances, molecular assays are being increasingly adopted by clinical laboratories to complement conventional methods, providing new diagnostic capabilities. L. Putignani Microbiology Unit, Children’s Hospital, Healthcare and Research Institute Bambino Gesù, Piazza Sant’Onofrio 4, 00165 Rome, Italy S. D’Arezzo and M. G. Paglia National Institute for Infectious Diseases “Lazzaro Spallanzani” I.R.C.C.S, Via Portuense 292, 00149 Rome, Italy P. Visca National Institute for Infectious Diseases “Lazzaro Spallanzani” I.R.C.C.S, Via Portuense 292, 00149 Rome, Italy Department of Biology, University of Roma Tre, Viale Marconi 446, 00146 Rome, Italy e-mail: visca@uniroma3.it Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_17, # Springer-Verlag Berlin Heidelberg 2010 357 358 17.1 L. Putignani et al. Introduction Fungal pathogens are currently rivaling their bacterial counterparts as emerging agents of nosocomial and community-acquired infections, posing a particular risk to patients under sustained immunosuppression. Deep mycoses and systemic mycotic infections have gained importance as potentially life-threatening opportunistic infections. This is due to the increasing number of immunocompromised individuals or the underlying risk factors, such as AIDS and cancer patients, organ transplant recipients, and patients under intensive care or predisposing concomitant treatments such as steroid, immunosuppressive, antineoplastic, and antibiotic chemotherapy. Indeed, the progress in transplant medicine and the therapy of hematological malignancies are counteracted by the threat of invasive or disseminated fungal infections (IFI), mainly triggered by species belonging to genus Candida and Aspergillus (Kontoyiannis and Bodey 2002; Pfaller and Diekema 2007). However, cutaneous fungi are also becoming emerging agents of hematogenously disseminated infections in the immunosuppressed host with severe or fatal prognosis (Cornely 2008). Recently, a substantial progress in fungal classification and identification has been achieved by means of molecular studies aimed at (1) tracing of evolutionary link between groups at higher taxonomic ranks through phylogenetic investigation; (2) improving taxonomy, mostly at the level of genera and species; (3) developing diagnostic applications for new defined taxonomic units; and (4) improving the epidemiological tools to monitor outbreaks and transmission routes of infection to subspecific entities (Yeo and Wong 2002). These broad aims have been achieved by optimizing techniques (Guarro et al. 1999) based on highly performing molecular targets, which serve as evolutionary clocks for phylogenetic studies (Yeo and Wong, 2002). One such target is the group of genes encoding the nuclear ribosomal RNA (rRNA). The main reasons for the success of the ribosomal DNA (rDNA) as an evolutionary marker is that its sequences encode for multiple-copy loci, whose repeated copies in tandem are synchronized by concerted evolution, and it is therefore reasonably treated as a single locus (Guarro et al. 1999). Furthermore, ribosomes are present in all organisms, with a common evolutionary origin. Parts of the molecule are highly conserved (van de Peer et al. 1996; van de Peer et al. 1997) and serve as reference points for evolutionary divergence studies. The 5.8S, 18S, and 25–28S rRNA genes (or rDNAs) are transcribed as a 35S to 40S precursors, along with internal and external transcribed spacers (ITS and ETS, respectively), which are spliced out of the transcript (Guarro et al. 1999) (Fig. 17.1). The conserved regions alternate with divergent domains (D1 and D2) and highly variable regions (ITS) (Hassouna et al. 1984) (Fig. 17.1). Between each cluster, there is a nontranscribed or intergenic spacer (NTS or IGS) that serves to separate the repeats from one another along the chromosome (Fig. 17.1). A 5S rRNA gene takes a variable position and transcription direction depending on the fungal group (Fig. 17.1). The total length of one DNA repeat is between 7.7 and 24 Kb (Hibbett 1992). For phylogeny of filamentous fungi, the 18S rDNA (also called 17 DNA-Based Detection of Human Pathogenic Fungi 359 Fig. 17.1 Schematic drawing of the physical organization of the fungal ribosomal rRNA gene cluster. All cluster components are included in the representation. SSU, small subunit (18S rRNA gene); LSU, large subunit (25–28S rRNA gene); D1 and D2 regions are the highly divergent regions of the LSU rRNA gene, shown as black boxes. ITS1, ITS2, and ETS1, ETS2 represent, respectively, intergenic and extragenic spacers, shown as dark grey boxes. Intergenic spacers (IGS1 and IGS2), between each cluster are nontranscribed regions which separate rRNA clusters from one another along the chromosome the small-subunit, SSU) is most often used either as full-length sequence or as subdomain of ca. 600 bp (Bruns et al. 1992). The divergent domains of the 25–28S (also called the large-subunit, LSU) rDNA are very informative and allow comparisons from high taxonomic levels down to the species level, although only a limited number of variable positions are present (Guého et al. 1993). In the 18S rRNA gene, the variable domains mostly provide insufficient information for diagnostic purposes (de Hoog and Gerrits van den Ende 1998), and large parts of the molecule must be sequenced to obtain the resolution required for species identification (de Hoog and Gerrits van den Ende 1992). In contrast, the 5.8S rDNA is too small and has the least variability. The 5S molecule has mainly been used to infer relationships at the order level, where differences could be traced back to the secondary structure of the molecule (Walker and Doolittle 1982). The ITS regions are much more variable, and sequences can be aligned with confidence only between closely related taxa. These regions have been extensively used to assess intraspecies differentiation (Kurtzman and Robnett 1991, 1998; Kurtzman and Fell 1998; Lott et al. 1998; Iwen 2003; Hinrikson et al. 2005). In yeasts, the D1 and D2 variable regions of the 25–28S rDNA have been extensively used for species-level identification (Sandhu et al. 1995; Sanguinetti et al. 2007; Linton et al. 2007; Putignani et al. 2008b) (Fig. 17.1). A region encompassing the D2 domain has also been exploited to produce a commercial sequencing kit based on the interrogation of libraries of fungal D2 rDNA sequences (MicroSeq D2 LSU rDNA Fungal Identification Kit, Applied Biosystems). The MicroSeq system is composed of PCR and cycle sequencing modules, identification and analysis software, and a D2 sequence library (Hall et al. 2003, 2004). Also mitochondrial targets, considered as “multicopy” loci because of the multiplicity of mitochondria per cell, have been exploited as molecular tools for the classification and identification of molds and yeasts (Guarro et al. 1999; Wang et al. 2000; Yeo and Wong 2002; Yamada et al. 2004). Introns of several protein-encoding genes, such as the b-tubulin (Tsai et al. 1994), actin (Cox et al. 1995), chitin synthase (Bowen et al. 1992), acetyl coenzyme A synthase (Birch et al. 1992), glyceraldehyde-3-phosphate dehydrogenase (Harmsen et al. 1992), lignin peroxidase (Naidu et al. 1990), or orotidine 50 -monophosphate decarboxylase (Radford 1993) genes, can also provide information at the species level. Recently, repetitive genome 360 L. Putignani et al. sequences (rep) have also been exploited as taxonomical tools for identification of Aspergillus spp. (Healy et al. 2004; Hansen et al. 2008) and Candida spp. (Wise et al. 2007) or for genomic fingerprinting (rep-PCR) assays adapted to an automated platform (DiversiLab system, Biomerieux). Amongst clinically relevant fungi, there are three main groups quite different from one another and classified according to their biological characteristics. One group is composed of dimorphic saprobes, which include soil-borne fungi that have developed the ability to switch from a yeast to a hyphal morphology to adapt to the hostile environment of the human body (Guarro et al. 1999). The dimorphic pathogens (e.g., Histoplasma capsulatum, Coccidioides immitis, Blastomyces dermatitidis, Paracoccidioides brasiliensis, Sporothrix schenckii, Penicillium marneffei) are incorporated in the Ascomycota and belong to the Plectomycetes group, a class of the supraordinal systematics defined according to the asci arrangement (Müller and von Arx 1973). The second group, which is the most numerous, consists of opportunistic saprobes, which cause opportunistic mycoses in individuals whose immune system is deficient or artificially suppressed (Kendrick 1992). This group includes Aspergillus, Fusarium, Rhizopus and Mucor genera, dematiaceous fungi, yeasts (Candida, Cryptococcus, Trichosporon), and zygomycetes. The third large group refers to dermatophytes, a group of obligate parasites, which attack the human skin, nails, and hair and are therefore mainly related to superficial mycoses (Kendrick 1992). This group includes Epidermophyton, Trichophyton, and Microsporum genera. From a taxonomical view point, the pathogenic and opportunistic fungi are distributed among three major phyla of the Kingdom Fungi: Ascomycota, Basidiomycota, and Zygomycota (Guarro et al. 1999). Recently, a new comprehensive phylogenetic classification of the Kingdom Fungi has been proposed, with reference to recent molecular phylogenetic analyses, and supported by several members of the fungal taxonomic community (Hibbett et al. 2007). Subkingdoms Dikarya and Basal Fungi include the main phyla of medical relevance. The dermatophytes are not a particular phylum but rather a short-hand label for the group of the three fungal genera Epidermophyton, Trichophyton, and Microsporum, all belonging to Ascomycota phylum (Fig. 17.2). Generally, the large fungal Phylum Glomeromycota are plant symbionts, Chytridiomycota and Neocallimastigomycota are animal pathogens, while Blastocladiomycota are algal pathogens (Hibbett et al. 2007) (Fig. 17.2). Apart from the taxonomic reevaluation, the teleomorph or anamorph relationship (sexual and asexual reproductive stages, respectively) within the phyla Ascomycota and Basidiomycota must also be taken into account for correct identification and proper description of medically important fungi, especially in laboratory reports to the clinical staff. Therefore, the species names reported hereafter will usually refer to the anamomorph nomenclature or, alternatively, to the teleomorph name when commonly used in practice. However, nomenclature interconversion from teleomorph to anamorph can be promptly achieved by interrogation of dedicated internet-based taxonomy browsers (e.g., http://www.doctorfungus.org/imageban/help. htm; http://www.ncbi.nlm.-nih.gov/Taxonomy/Browser/wwwtax.cgi). 17 DNA-Based Detection of Human Pathogenic Fungi 361 Fig. 17.2 Schematic phylogeny and classification of fungal human pathogens within the Basal Fungi and Dikarya Subkingdoms. The pathogenic and opportunistic human fungi are distributed among the three major phyla Ascomycota, Basidiomycota, and Zygomycota of the Kingdom Fungi. Subkingdom Dikarya include both Ascomycota and Basidiomycota (dark grey), while Zygomycota (pale grey) belong to the Basal Fungi despite a pending resolution of relationships among clades that still affect its definitive taxonomic location. The phylum Glomeromycota includes important plant symbionts; Chytridiomycota and Neocallimastigomycota animal pathogens, while Blastocladiomycota algal pathogens. The group of dermatophytes belong to the Ascomycota phylum. The branch lengths are not proportional to genetic distances. Modified from Hibbett et al. (2007) From a clinical perspective, early and accurate diagnosis of fungal infections is crucial to avoid the extensive clinical use of empirical antifungal therapy, which is the primary cause for the emergence of antifungal resistance (Yeo and Wong 2002). It must be pointed out that a major drawback to the successful treatment of IFIs is the lack of sensitive and specific methods for early diagnosis. Standard approaches to the laboratory diagnosis of IFIs include (1) direct microscopic visualization for the presence of organisms in freshly obtained body fluids, (2) detection of specific antibodies, (3) histopathologic demonstration of fungi within tissue sections, and (4) classical cultivation of the causative fungus with its subsequent macroscopic and biochemical identification. These approaches are not sensitive enough and/or 362 L. Putignani et al. specific to the diagnosis of IFI, which requires invasive procedures to obtain the necessary quantity of specimens. Moreover, the diagnosis of IFI can be biased by improper samplings, not representative of the real agent and site of the infection, as well known for invasive pulmonary aspergillosis (Kappe and Rimek 1999). Furthermore, phenotypic identification of fungi requires complicated algorithms and time-consuming procedures, not always correctly interpretable in daily diagnostic routine without dedicated mycologists. Among the culture-independent methods, detection of a specific host antibody response is attractive because such tests can be performed rapidly and do not require invasive sampling procedures. However, presence of host antibodies does not always correlate with the presence of invasive mycosis, especially in patients whose humoral response is impeded by immunosuppressive drugs and/or serious underlying disease. Detection of macromolecular antigens generally requires a relatively large fungal burden, which may limit the sensitivity of these assays. Nonetheless, several examples of successful antigen detection systems exist, and some of these are widely used in the clinical mycology laboratory. Alternatives to standard culture and serologic diagnostic methods include amplification and detection of specific fungal DNA sequences and the detection and quantitation of specific fungal metabolome and proteome products. In this scenario, the diagnosis of IFIs remains a challenge because clinical symptoms are not pathognomonic, and searching for mycotic agents is delayed unless a high index of clinical suspicion is applied and a differential diagnosis provided. This chapter will address recent advances in the DNA-based diagnosis of relevant or emerging fungal pathogens, with special attention to the IFIs, which represent an important cause of morbidity and mortality in both the developed and developing world. Progresses in advanced molecular methods for diagnosis and epidemiological typing of pathogenic fungi are becoming fundamental for early treatment of patients, controlling fungal clearance, and counteracting resistance to antifungal therapy. 17.2 DNA Manipulations Specimen handling and preparation have a significant impact on the performance of molecular diagnostic tests for fungal detection. The sample preparation method should release intracellular DNA from the fungal cell wall, concentrate DNA targets that may be present in very small amounts, and eliminate contaminants, potential inhibitors, and other extraneous materials without degrading the target DNA. The availability of an easy-to-perform DNA extraction procedure, providing pure DNA devoid of PCR inhibitors, would be ideal for any PCR-based diagnostic test. Simple cytolytic procedures for DNA extraction, e.g., thermolysis, although used in some protocols (Putignani et al. 2008a, b) cannot be applied to all fungi. For instance, filamentous fungi have strong cell walls which are often resistant to traditional DNA extraction procedures. Fungal nucleases, polysaccharides, and 17 DNA-Based Detection of Human Pathogenic Fungi 363 pigments also contribute to difficulties in purifying DNA from filamentous fungi (Hope et al. 2005). There are a multitude of nucleic acids extraction techniques (Griffiths et al. 2006). The preferable method represents a compromise between efficiency, purification yields, and transferability to the laboratory routine. DNA may be extracted using in-house methods, commercial kits, and automated commercial techniques. Mechanical destruction with glass beads and freeze–thaw steps with liquid nitrogen or a heat-alkali treatment have successfully been applied (Hopfer et al. 1993; Löffler et al. 1997; Griffin et al. 2002). DNA extraction following enzymatic digestion of the fungal wall is another effective method (Williamson et al. 2000). However, many of these in house-methods are not suited for the clinical microbiology laboratory, where many samples are simultaneously processed. Moreover, the use of toxic chemicals such as phenol–chloroform mixtures further limits the use of in house-methods in the clinical routine (Griffiths et al. 2006). The use of commercial kits (e.g., QIAmp Tissue, Qiagen; GeneReleaser, BioVentures; Puregene D 6000, Gentra; Dynabeads DNA DIRECT, Dynal; and DNAzol, Molecular Research Center) shortens the extraction procedure, but the efficiency of extraction of fungal DNA can vary considerably between commercial kits (Griffiths et al. 2006; Löffler et al. 1997). Automated commercial techniques (e.g., MagNA Pure LC; Roche Diagnostics) are better suited for routine clinical laboratories (Costa et al. 2002). High-speed cell disruption (HSCD) incorporating chaotropic reagents and lysing matrices provides rapid lysis of cells and high yields of DNA from medically important yeasts (e.g., Candida albicans, Cryptococcus neoformans, Trichosporon beigelii) and filamentous fungi (e.g., Aspergillus spp. and Fusarium solani) (Müller et al. 1998). Concerted efforts are focused on the optimization of DNA extraction methods from yeasts, particularly from blood samples in cases of candidemia (Metwally et al. 2008a, b). Although the quality of samples can affect the recovery of nucleic acids (Bougnoux et al. 1999; Fredricks et al. 2005; Metwally et al. 2008a, b bis), yeast DNA has successfully been extracted and purified from different clinical samples, including whole blood (Buchman et al. 1990), serum and plasma (Kan 1993; Bougnoux et al. 1999; Metwally et al. 2008a, b, bis), bronchoalveolar lavage fluid (Klingspor and Jalal 2006), and cerebrospinal fluid (CSF) (Ralph and Hussain 1996), using a variety of home-made and commercial kits protocols. Sample contamination should also be considered as a major problem in molecular diagnosis because of the extremely high sensitivity of all nucleic acids amplification techniques. Fungal spores, such as conidia from Aspergillus spp. and other molds, might be present in the air. Thus, airborne spore inoculation during the DNA extraction process could potentially lead to false-positive results, especially if panfungal primers are applied (Löffler et al. 1999). However, the risk of contamination is not higher in fungal PCR assays than in other diagnostic PCRs if general precautions are taken. In order to control naturally arising DNA from airborne sources (e.g., fungal spores) negative controls should be included during each DNA extraction procedure. Negative controls consist of sterile water or blood from healthy individuals, and should be subjected to all preparation steps in parallel with the extracted samples (Sarkar and Sommer 1990; Löffler et al. 1999). 364 L. Putignani et al. The development of semi-automated platforms that comprise nucleic acids extraction and product detection through a series of linked instruments have produced substantial implementation of molecular testing within the routine of the clinical diagnostic laboratory, strongly reducing the risk of sample contamination. The absence of post-PCR processing after amplification step, as for the on-chip platforms and the real-time PCR-based procedures, has offered many practical advantages over the use of traditional detection methods by facilitating sample processing and minimizing DNA shedding in the laboratory environment (Löffler et al. 1999). 17.3 Panfungal Assays For clinical diagnostic purposes, the broad-range detection of pathogenic fungi in clinical samples is as important as the ability to identify the specific pathogen(s). The common approach involves the application of broad-ranging panfungal primers with postamplification analysis for species determination (Table 17.1). Panfungal primers are directed toward conserved regions, usually within multicopy genes, which flank sequences containing species specific polymorphisms that can be defined in postamplification analysis (Chen et al. 2002). Depending on the assay, a diverse range of fungal genera and species can be identified, including species of Aspergillus, Candida, Cryptococcus, Fusarium, Trichosporon, Rhizopus spp., etc. (Makimura et al. 1994; van Burik et al. 1998; Hendolin et al. 2000; Klingspor and Jalal, 2006; Lau et al. 2007; Schabereiter-Gurtner et al. 2007; Spiess et al. 2007; Zeng et al. 2007) (summarized in Table 17.1). The multicopy ribosomal gene complex is a useful target for these assays for reasons of sensitivity (multicopy target), high sequence conservation of 5.8S, 18S, and 25–28S rDNA regions (for panfungal primers), and high variability of its intervening ITS regions (for species specific probes) with high interspecies and low intraspecies heterogeneity (Lott et al. 1998; Iwen 2003; Hinrikson et al. 2005). Panfungal assays are potentially able to detect all fungal pathogens, but require additional tests for specieslevel identification. The most common approaches for species differentiation rely upon differences in restriction enzyme digestion patterns of amplicons, or their hybridization with species-specific probes (Hopfer et al. 1993; Sandhu et al. 1995; Einsele et al. 1997; Elie et al. 1998; Martin et al. 2000). Other methods, such as differences in PCR product sizes following electrophoresis, amplicon sequencing, or single-stranded conformational polymorphism analysis, have been applied (Walsh et al. 1995; Henry et al. 2000; Chen et al. 2002; Iwen 2003; Gupta et al. 2004; Iwen et al. 2004). Detection of amplicons can also be achieved by means of hybridization with enzyme-labeled oligonucleotide probe, eventually in microtiter plate-based enzyme immunoassay (Elie et al. 1998; Löffler et al. 1998; Wahyuningsih et al. 2000), or fluorogenic probes (Guiver et al. 2001; Maaroufi et al. 2003; Selvarangan et al. 2003), or by Southern blotting (Sandhu et al. 1995; Einsele et al. 1997; van Burik et al. 1998; Evertsson et al. 2000). The use of a 17 Specificity References Einsele et al. (1997) Walsh et al. (1995) 18S rRNA 18S rRNA Blood Culture 28S rRNA Cervical swab, nail and Real Time horny skin scraping, serum, blood, urine 28S rRNA Blood PCR Southern blot 28S rRNA Respiratory (BAL) and tissues Blood, respiratory, tissue Skin, nail, wound, urine, blood, respiratory, tissue Fresh and formalinfixed, paraffinembedded tissue Real Time FRET Real Time TaqMan Aspergillus spp., Candida spp. Aspergillus spp., Candida spp., Cryptococcus neoformans, Pseudallescheria boydii, Rhizopus arrhizus Aspergillus spp., Candida spp. Absidia spp., Mucor spp., Rhizopus spp., Rhizomucor spp. Aspergillus spp., Candida spp., Cryptococcus spp., Mucor spp., Penicillium spp., Pichia spp., Microsporum spp., Trichophyton spp., Scopulariopsis spp. Aspergillus spp., Candida spp., C. neoformans Rhizopus spp., Mucor spp., Rhizomucor spp. Aspergillus spp., Candida spp. PCR ELISA Aspergillus spp.,Candida spp. Badiee et al. (2007) PCR Sequencing Lau et al. (2007) PCR Multiplex liquid hybridization and sequencing Candida spp., Cryptococcus spp., Trichosporon spp., Aspergillus spp., Fusarium spp., Scedosporium spp., Exophiala spp., Exserohilum spp., Apophysomyces spp., Actinomucor spp., Rhizopus spp. Aspergillus spp., Candida spp., C. neoformans 28S rRNA ITS ITS1 Southern blot Ethidium bromide staining TaqMan and sequencing Van Burik et al. (1998) Machouart et al. (2006) Vollmer et al. (2008) Evertsson et al. (2000) Kasai et al. (2008) Basková et al. (2007) Hendolin et al. (2000) (continued) 365 ITS1-5.8S Tissues rRNA-ITS2 PCR PCR-RFLP DNA-Based Detection of Human Pathogenic Fungi Table 17.1 Main DNA-based wide-broad range and panfungal assays Target DNA Method Detection method Specimen 18S rRNA Blood PCR Southern blot Ethidium bromide 18S rRNA Culture PCR-SSCPa staining Method Detection method Specificity References INNO-LiPA Enzyme immunoassay Microarray hybridization Aspergillus spp., Candida spp., Cryptococcus spp. Aspergillus spp., Candida spp., Fusarium spp., Mucor racemosus, Rhizopus microsporus, Scedosporium prolificans, Trichosporon asahii Aspergillus spp., Candida spp., Epidermophiton floccosum, Microsporum spp., Trichophiton spp. Aspergillus spp., Candida spp., Cryptococcus spp. Martin et al. (2000) Multiplex-PCR ITS2 Culture PCR ITS2 Culture Repetitive sequences Cytochrome b Culture Reverse line blot hybridization Rep-PCR Microfluidics chip Culture and tissues Real Time Spiess et al. (2007) Turenne et al. (1999) Playford et al. (2006) TaqMan Coccidioides spp., Blastomyces Pounder et al. (2006) dermatitidis, Histoplasma capsulatum Hata et al. (2008) Absidia spp., Apophysomyces spp., Cunninghamella spp., Mucor spp, Rhizopus spp, Saksenaea spp. Aspergillus spp., Candida spp. Klingspor and Jalal (2006) Melting point Aspergillus spp., Candida spp. FRET Chemiluminescence Aspergillus spp., Candida spp., Cryptococcus spp. Schabereiter-Gurtner et al. (2007) Zeng et al. (2007) L. Putignani et al. Blood, respiratory, bile, Real Time drainage, urine, pleura, CSF, biopsy ITS2 Blood, respiratory, Real time tissues ITS1/ITS2 Blood, respiratory PCR-reverse line blot (BAL), tissue, CSF, (RLB) skin a Single-strand conformational polymorphism 18S rRNA Fluorescent capillary electrophoresis Enzyme immunoassay 366 Table 17.1 (continued) Target DNA Specimen ITS1-5.8S Culture rRNA-ITS2 18S rRNA, Blood, tissues and 5.8S, ITS1 respiratory (BAL) 17 DNA-Based Detection of Human Pathogenic Fungi 367 panfungal PCR followed by hybridization with species-specific probes is a practical solution to the problem of fungal detection. 17.4 Aspergillus spp. Aspergillus spp. are filamentous, cosmopolitan, ubiquitous fungi which can cause life-threatening infections, especially in immunocompromised patients. Aspergillus spp. are commonly isolated from the soil, plant debris, and the indoor environment, including hospitals (Latgé 1999). The genus Aspergillus includes over 185 species. Nearly 20 species have been reported as causative agents of opportunistic infections in humans. A. fumigatus is the most frequently isolated species in the clinical setting. Other common species associated with infection are A. flavus, A. niger, and A. terreus, while A. nidulans, A. versicolor, A. candidus, A. oryzae, A. sydowii, and A. clavatus have been rarely documented (Patterson 2005). Aspergillosis includes a large spectrum of fungal diseases which primarily affect the lung. The transmission of fungal spores to the human host is via inhalation (Zmeili and Soubani 2007). Aspergillus spp. may cause a variety of pulmonary diseases, depending on immune status and the presence of underlying lung disease. These manifestations range from hypersensitivity reactions (allergic bronchopulmunary aspergillosis, ABPA) to noninvasive colonization of previously damaged tissue (pulmonary aspergilloma) to acute or chronic limited invasive disease (chronic necrotizing pulmonary aspergillosis) to rapidly progressive invasive disease (invasive aspergillosis) (Patterson 2005; Zmeili and Soubani 2007). Invasive aspergillosis is an often fatal infection that occurs in severely immunosuppressed patients, and is characterized by invasion of blood vessels and organ dissemination, resulting in significant morbidity and mortality (Kontoyiannis et al. 2002). Risk factors for invasive aspergillosis are severe neutropenia, hematopoietic stem cell and solid organ transplantation, prolonged and high-dose corticosteroid therapy, hematological malignancy, cytotoxic therapy, AIDS, and chronic granulomatous disease (Segal and Walsh 2006; Zmeili and Soubani 2007). The incidence of life threatening invasive Aspergillus infections has been increasing with the growing number of transplant patients and patients with leukemia, lymphoma, and other malignancies. Incidence rates of invasive Aspergillus infection are about 17–26% in lung transplant patients, 5–24% in acute leukemia patients, 5–15% in allogenic bone marrow transplant patients, 2–13% in heart transplant patients, and 1–3% in lymphoma patients (Patterson et al. 2000; Kontoyiannis et al. 2002). 17.4.1 Laboratory Diagnosis Given that invasive Aspergillus infections are associated with high crude and attributable mortality rates, early, rapid, and accurate diagnosis is important in 368 L. Putignani et al. order to guide the selection of appropriate antifungal therapy and thus improve patient outcomes, as well as for epidemiological purposes. Earlier detection of infection permits prompt initiation of antifungal therapy with greater likelihood for improved survival and reduced morbidity. Moreover, diagnostic tests with a high negative predictive value may allow expensive and potentially toxic antifungal drugs to be withheld (Hope et al. 2005). The traditional diagnostics include microscopy, culture on agar media, antigenemia, and search for antibodies. All of them are limited either by poor sensitivity, narrow temporal window for fungal detection, complex interpretation and high levels of nonspecific reactions. Positive culture of Aspergillus spp., although indicative, is not a substantial proof of infection. Furthermore, culture-based phenotypic identification techniques are slow and prone to misidentification (Reiss and Morrison 1993). The “gold standard” investigation remains microbiological and/or histological evidence of tissue invasion, which is not always achievable (Bretagne and Costa 2005). The detection of galactomannan has been included into the routine practise for diagnosis of invasive aspergillosis (Hope et al. 2005). There are two commercial assays for the detection of galactomannan: the latex agglutination test (e.g., Pastorex Aspergillus; Sanofi Diagnostics Pasteur) and a double-antibody sandwich enzyme immunoassay (e.g., Bio-Rad Platelia1 Aspergillus EIA, Bio-Rad Laboratories). However, the growth phase, microenvironment, host immune status, and pathology may all influence galactomannan release and hence the results of immunological tests (Latgé et al. 1994). Furthermore, cross reactivity with other filamentous fungi, bacteria, drugs, and cotton swabs have been documented (Kappe and Schulze-Berge 1993; Hashiguchi et al. 1994; Swanink et al. 1997; Dalle et al. 2002; Mennink-Kersten et al. 2004). 17.4.2 Molecular Detection The application of PCR technology to molecular diagnostics holds great promise for the early identification of Aspergillus spp. (Williamson and Leeming 1999; Klingspor and Loeffler 2009). DNA-based assays rapidly detect the presence of fungal DNA in blood and other sterile body fluids with high sensitivity and specificity (Williamson and Leeming 1999), particularly since parts of the fungal genome, especially multicopy gene targets, were identified and sequenced (Buchheidt and Hummel 2005). For early detection of Aspergillus spp. in clinical samples, several groups have succeeded in defining target gene sequences, practicable primers, and effective DNA extraction methods. Technical difficulties, such as differentiation and specification of the amplicons, have also been overcome by different approaches (e.g., species-specific oligonucleotide probes and PCR ELISA). A variety of PCR protocols for human samples have been published, including panfungal PCR assays (Table 17.1) and methods that detect one species or genus (Table 17.2). Using different primer sets, PCR has been recommended by several studies as a useful tool in establishing the diagnosis of invasive aspergillosis Culture PCR-SSCP ITS1-5.8S rRNA-ITS2 Culture PCR ITS2 Culture and tissue Blood Culture PCR Blood culture Real Time PCR Real-Time Septifast References Aspergillus spp. A. fumigatus Aspergillus spp. Makimura et al. (1994) Löffler et al. (2000) Kami et al. (2001) A. fumigatus, A. flavus, A. glaucus, A. niger, A. terreus A. fumigatus, A. flavus, A. nidulans, A. niger, A. ochraceus, A. terreus, A. ustus, A. versicolor A. fumigatus, A. flavus, A. terreus, A. nidulans, A. niger Aspergillus spp. A. fumigatus A. fumigatus Sanguinetti et al. (2003) A. fumigatus, A. flavus, A. terreus, A. niger, A. ustus, A. nidulans A. fumigatus, A. flavus, A. terreus, A. niger, A. nidulans A. fumigatus Henry et al. (2000) A. fumigatus, A. flavus, A. terreus, A. niger, A. nidulans, A. ustus, A. versicolor A. fumigatus, A. flavus, A. terreus, A. niger A. fumigatus, A. flavus, A. terreus, A. niger De Aguirre et al. (2004) TaqMan Ethidium bromide staining Melting curve A. fumigatus Halliday et al. (2005) Melchers et al. (1994) Pham et al. (2003) Challier et al. (2004) Spreadbury et al. (1993) Rath and Ansorg (2000) Zhao et al. (2001) Faber et al. (2009) Logotheti et al. (2009) Mancini et al. (2008) (continued) 369 18S rRNA ITS1-5.8S rRNA-ITS2 and aspergillopepsin 1st and 4th exon ITS Silver staining Ethidium bromide staining ELISA Specificity DNA-Based Detection of Human Pathogenic Fungi ITS1-5.8S rRNA-ITS2 17 Table 17.2 Main DNA-based assays for Aspergillus spp. identification Target DNA Specimen Method Detection method 18S rRNA Respiratory PCR Southern blot 18S rRNA Blood Real Time FRET 18S rRNA Blood and Real Time TaqMan plasma 18S rRNA Respiratory Real Time TaqMan (BAL) Real Time TaqMan 18S rRNA Blood, CSF, ascitic fluid, tissue 18S rRNA Respiratory PCR-RFLP Southern blot (BAL) 5.8S rRNA Serum Real Time TaqMan 28S rRNA Serum Real Time TaqMan 26S-ITS Respiratory PCR Southern blot (BAL) ITS1-5.8S rRNA-ITS2 Culture PCR Sequencing Specimen Mitochondrial tRNA Respiratory (BAL) Mitochondrial tRNA Respiratory (BAL) Mitochondrial tRNA Tissues Mitochondrial tRNA Blood and serum Mitochondrial tRNA Serum Mitochondrial tRNA Respiratory (BAL) and tissues Mitochondrial cytochrome b Respiratory (BAL) and blood IgE-binding protein Blood and urine FKS Blood and serum Repetitive sequences Culture Repetitive sequences Culture Method PCR Detection Specificity method Southern blot A. fumigatus, A. flavus, A. terreus, A. niger Bretagne et al. (1995) PCR Southern blot A. fumigatus, A. flavus, A. terreus, A. niger Raad et al. (2002) PCR Real Time Sequencing TaqMan Aspergillus spp. A. fumigatus Rickerts et al. (2006) Costa et al. (2001) Real Time Real Time FRET FRET A. fumigatu, A. flavus A. fumigatus Costa et al. (2002) Rantakokko-Jalava et al. (2003) Real Time FRET A. fumigatus Spiess et al. (2003) PCR Southern blot A. fumigatus Reddy et al. (1993) Real Time TaqMan Rep-PCR Microfluidics A. fumigatus, A. flavus, A. terreus chip Microfluidics A. fumigatus, A. flavus, A. terreus, A. niger, chip A. ochraceus, A, candidus, A. sydowii, A. versicolor, A. nidulans Rep-PCR A. fumigatus References 370 Table 17.2 (continued) Target DNA Costa et al. (2001) Healy et al. (2004) Hansen et al. (2008) L. Putignani et al. 17 DNA-Based Detection of Human Pathogenic Fungi 371 (Spreadbury et al. 1993; Tang et al. 1993; Melchers et al. 1994; Skladny et al. 1999). However, the lack of standardization of technical issues continues to represent a considerable barrier for the widespread application of PCR in the diagnosis of invasive aspergillosis (Bretagne 2003; Hope et al. 2005; Mengoli et al. 2009). PCR assays may be applied to broncho alveolar lavage (BAL) specimens, blood fractions (serum, plasma, whole blood) and tissues, including paraffinembedded thin sections (Hope et al. 2005). Initial studies focused on detection of DNA in BAL samples (Spreadbury et al. 1993; Makimura et al. 1994; Melchers et al. 1994; Bretagne et al. 1995; Raad et al. 2002) (Table 17.2). With few exceptions, the rate of positivity was attained in 35% of the samples (Bretagne 2003). However, the inhaled Aspergillus spores or conidia, which are ubiquitous in the air, could lead to false-positive reactions. In fact, Aspergillus hyphae are often present in the air and could transiently colonize the respiratory tract of noninfected individuals or even contaminate the amplification mixture. The PCR performed on BAL cannot differentiate between infecting and colonizing fungi in the oropharyngeal and bronchoalveolar space of patients (Bretagne and Costa 2005). On the other hand, some authors have highlighted the potential value of a negative result to exclude a diagnosis of aspergillosis (Raad et al. 2002). Therefore, a negative PCR test result in a patient with a suspected infection suggests that the patient does not have invasive pulmonary aspergillosis, and most likely does not have organism colonization. Given the high rate of positivity and the difficulties in interpreting PCR positive results in bronchial specimens, several studies have switched to blood samples, reporting a high positive predictive value (>95%) (Reddy et al. 1993). The use of PCR technology with serum or plasma has several advantages over the use of BAL samples. First, assuming appropriate handling of the specimen, false-positive results do not occur from specimen contamination. Second, taking blood sample is considerably easier than obtaining BAL (Einsele et al. 1997). Compared to ELISA, however, PCR positivity seems to occur later than galactomannan detection (Bretagne et al. 1998). Therefore, the combined use of PCR and ELISA should result in a definitive diagnosis aspergillosis, even in the absence of obvious clinical signs (Bretagne et al. 1998). The optimal blood fraction for the detection of Aspergillus DNA remains unknown. One study, using quantitative PCR (qPCR), suggested that the yield of DNA from serum, plasma, and white cells was similar (Costa et al. 2002), while another demonstrated that the PCR signal from whole blood was superior to that from plasma (Löffler et al. 2000). Serum has the advantage of enabling concomitant antigen testing (Costa et al. 2002), and does not require the addition of anticoagulants (e.g., sodium citrate, edetic acid, or heparin) that may inhibit PCR. An important study has recently focused on the diagnostic applicability of serial blood and serum samples to diagnose IA in neutropenic patients by means of real-time quantitative PCR combined with galactomannan quantification (Cuenca-Estrella et al. 2009). Furthermore, a previous report (Suarez et al. 2008) had shown that real-time PCR assays, performed by using DNA extracted from large serum volumes, may actually represent a robust and early diagnostic tool for IA in patients under hematologic surveillance. 372 L. Putignani et al. A variety of in-house and commercial PCR-based assays have been developed for the identification of Aspergillus species (Table 17.2). The mitochondrial tRNA genes and the (apo)cytochrome b have been used as PCR targets (Wang et al. 2000) (Table 17.2). The rep-PCR assay (DiversiLab Aspergillus system) enables rapid and accurate Aspergillus spp. identification (Healy et al. 2004; Hansen et al. 2008). This approach takes advantage of the fact that there are repetitive elements interspersed throughout the fungal genome that, when amplified by PCR, give highly discriminatory reproducible profiles within Aspergillus spp. The method has been developed as a user-friendly kit, and the automated detection and analysis provides readily interpretable reports (Hansen et al. 2008). Moreover, rep-PCR based identification is in full agreement with the ITS region sequence-based identification (Healy et al. 2004). The MicroSeq D2 LSU rDNA Fungal Identification Kit (Applied Biosystems) has allowed the identification of filamentous fungi, including Aspergillus spp., by exploiting the D2 LSU diversity region (Hall et al. 2004). Some of the newer assays are successfully using the real-time PCR technology, either the Light Cycler or TaqMan technology, both combining amplification with simultaneous amplicon detection (Table 17.2). The method applies to DNA extracted from both blood and BALs. One of the main advantages of the real-time PCR is the possibility of avoiding false positive results due to contamination with previously amplified products. Contamination is reduced because the reaction tubes need not be opened following amplification. Moreover, real-time PCR techniques can also include the systematic use of uracyl-N-glycosylase. The sensitive and specific quantification of the fungal burden seems to be of clinical relevance, since the assessment of the individual fungal burden may possibly allow therapeutic monitoring (Costa et al. 2001). To achieve an improved, specific, sensitive, and rapid method for quantification of the A. fumigatus fungal load in clinical samples, Spiess et al. (2003) established a LightCycler PCR assay to test blood and BAL samples. An optimal pair of primers and hybridization probes derived from the sequence of the A. fumigatus mitochondrial cytochrome b gene was selected (Spiess et al. 2003). Two new TaqMan-based PCR assays for a fungal species, one targeting a single copy gene and the other a mitochondrial gene have been developed (Costa et al. 2001). To diagnose invasive mold infection from serum samples a quantitative realtime PCR assay targeting the 5.8S rRNA gene was designed (Pham et al. 2003). This assay distinguishes invasive infections caused by Aspergillus spp. from those caused by Fusarium spp., and Scedosporium spp. (Pham et al. 2003). Recently, Vollmer et al. (2008) developed a broad-range 28S rDNA real-time PCR assay for the rapid detection of fungal pathogens in various clinical specimens (e.g., serum, urine, and EDTA-supplemented blood). The assay allows the simultaneous detection of and discrimination between genera of pathogenic fungi, including Aspergillus, Candida, Cryptococcus, Mucor, Penicillium, Pichia, Microsporum, Trichophyton, and Scopulariopsis (Vollmer et al. 2008). A rapid real-time PCR assay has recently been designed which exploits regions of the 18S rRNA locus to simultaneously detect the common A. fumigatus, A. flavus, A. terreus, and A. niger species by a differential melting point analysis (Faber et al. 2009). Concomitantly, the multiplex PCR assay described by Logotheti et al. (2009), has provided a similar panel for simultaneous 17 DNA-Based Detection of Human Pathogenic Fungi 373 identification (A. fumigatus, A. flavus, A. terreus, and A. niger) with an easier multiplexPCR system based on a combined ITS1-5.8S rRNA-ITS2 and aspergillopepsin first and fourth exon primer set. Recently, the real-time PCR assay based on ITS variability has been adapted to semiautomated platforms (Light Cycler SeptiFast, Roche) significantly reducing diagnostic turnaround time, particularly for the diagnosis of A. fumigatus, a typical slow-growing filamentous fungus (Mancini et al. 2008). 17.5 Blastomyces dermatitidis B. dermatitidis is a thermally dimorphic fungus and a probable saprobe of the soil. It is rarely isolated as a natural habitat, specifically inhabiting decaying wood material. Isolation from the environment is most likely when the sample contains soil and is rich in organic material. It is endemic in North America with highest incidence of infection in Mississippi, Ohio, and the Missouri valleys (Pappas 2004). African type B. dermatitidis strains isolated from cases in Africa also exist. It was demonstrated that African type strains are not identical to the North American strains. These two groups most probably constitute distinct types of B. dermatitidis showing geographic and serologic diversity (McCullough et al. 2000). The sexual state (teleomorph) of B. dermatitidis belongs to the family Onygenaceae and is referred to as Ajellomyces dermatitidis. B. dermatitidis is the only species included in the genus Blastomyces and is the causative agent of blastomycosis (Pappas 2004). Cutaneous and disseminated blastomycosis are the two clinical forms of the disease. Blastomycosis is generally acquired by inhalation, and initially presents with a pulmonary infection which may later disseminate to other organs. Primary cutaneous infection due to direct inoculation of the fungus into the skin is also likely. Hematogenous spread of the organism results in infection of skin, bones, kidneys, and male urogenital system. Reactivation blastomycosis and subclinical, self-limiting infections have been defined (Farr et al. 1992; Mounts and Deepe 1998). Although B. dermatitidis is a pathogenic fungus and blastomycosis occurs mainly in immunocompetent hosts, it may also infect immunocompromised patients, indicating that B. dermatitidis has now emerged as an opportunistic pathogen (Pappas 2004). 17.5.1 Laboratory Diagnosis The morphology of the fungus is mold-like at 25 C and yeast-like at 37 C. At 25 C, septate hyaline hyphae and unbranched short conidiophores are observed. In rich medium or in infected tissue sections the fungus appears as budding yeast cells. Conversion of the mold phase to the yeast phase for definitive identification is rarely performed now (Kauffman 2006). Diagnosis is often achieved by detection 374 L. Putignani et al. of specific antibodies, which are usually absent at presentation of symptoms, and which can be impaired in immunocompromised patients (Yeo and Wong 2002). 17.5.2 Molecular Detection Since therapy and prognosis depend on early and specific diagnosis, there has been considerable interest in molecular detection of B. dermatitidis. Specific DNA probes have successfully been used for identification of B. dermatitidis isolates through different hybridization assays (Stockman et al. 1993; Sandhu et al. 1995; Lindsley et al. 2001). A sensitivity of 95% and a specificity of 100% with B. dermatitidis yeast cells were obtained by using two oligonucleotide pairs complementary to the 18S and 28S rDNA (Hayden et al. 2001). A nested PCR assay was established targeting the gene encoding the species-specific B. dermatitidis adhesin (BAD), formerly called WI-1 (Bialek et al. 2003). This assay was specific and the detection limit of 0.1 fg target DNA was comparable to the 18S rDNA PCR (Bialek et al. 2005a, b). The 18S rDNA PCR is routinely used as a screening assay, and when positive, the specific nested PCR is added to confirm the diagnosis (Table 17.3). More recently, the automated rep-PCR (DiversiLab system) has been used to identify B. dermatitidis, Coccidioides, and H. capsulatum isolates, providing excellent performance for the identification of B. dermatitidis isolates, all of which had very similar fingerprint patterns (Pounder et al. 2006). 17.6 Candida spp. Candida is a complex genus comprising 163 anamorphic species with teleomorphs in at least 13 genera (Kurtzman and Fell 1998). Some of these genera, such as Pichia (Peterson and Kurtzman 1991) or Debaryomyces (Kurtzman and Robnett 1991, 1994) seem to be polyphyletic. Until few years ago, only a few pathogenic species of Candida were known, namely C. albicans, C. parapsilosis, Candida krusei, C. tropicalis, Candida lusitaniae, Candida dubliniensis, and Candida glabrata. However, in recent years the number of species related to human infections has increased considerably (Hazen 1995; D’Antonio et al. 1998; Sullivan and Coleman 1998; Trofa et al. 2008). Now nearly 20 species have been identified for being associated with human infection (Hazen 1996). Candida spp. can either colonize or infect nearly every body surface. The various forms of candidiasis are the most frequent causes of fungal infection in man, and can present with extremely diverse clinical manifestations. Candida spp. can produce infections in otherwise healthy individuals, as well as in individuals with impaired immune function. Candidiasis may be superficial (cutaneous), local (mucocutaneous, affecting mouth and vagina), deep-seated (involving central nervous system, respiratory and urinary tract, cardiac, ocular, peritoneum, and vasculature affections) and disseminated as 17 DNA-Based Detection of Human Pathogenic Fungi 375 systemic syndrome (candidemia), arisen from hematogenous spread from the primarily infected site. Systemic candidiasis is a complicate disease affecting individuals with reduced immune function or any other type of weakening of their defences. Almost any organ of the body may be involved, after beginning as an episode of candidemia, during which Candida can be isolated from blood. From a clinical standpoint, systemic candidiasis may differ into four forms: (1) the catheterrelated candidemia, due to infection of a vascular catheter; (2) the acute disseminated candidiasis in which candidemia is present and may apparently spread to one or more organs; (3) the chronic disseminated candidiasis or hepatosplenic candidiasis, occurring after prolonged episodes of bone marrow dysfunction and neutropenia that occur during treatment for leukemia; (4) the deep organ candidiasis, in which any organ may be affected, either alone or in combination. When Candida disseminates, multiple organs are usually involved, with the kidney, brain, myocardium, and eye being the most common. As an aid to earlier diagnosis, considerable attention has been focused on the detection of Candida antigens. Despite the appearance of a large number of reports on the serologic diagnosis of disseminated candidiasis, controversies remain regarding the value of various serodiagnostic procedures. Problems with the older diagnostic tests have been reviewed in detail (Edwards 1991). Among severely immunosuppressed patients, almost all patients with candidemia have disseminated disease. The problem is compounded by the absence of positive blood cultures in many patients with disseminated disease. In these cases, a positive blood culture for Candida cannot be underestimated but verified by repeated sampling. However, interpretation of the ensuing resultion of the candidemia should be made with the recognition that 50% of the patients with disseminated candidiasis would not have positive blood cultures especially when concomitant bacterimia exists (Hockey et al. 1982; Geha and Roberts 1994). Promptness in correct fungal identification is therefore crucial in laboratory diagnosis to overcome limits arising from both serological and culture tests. Candida infections are the most frequent cause of IFIs worldwide (Pfaller and Diekema 2007). In the United States, Candida spp. are the fourth most common cause of nosocomial bloodstream infection (Pfaller and Diekema 2007). C. albicans remains by far the most common species causing invasive candidiasis worldwide although the frequency of other species, including Candida tropicalis, Candida parapsilosis, Candida glabrata, and Candida krusei, has been steadily increasing over the last 10 years (Pfaller and Diekema 2007). The burden of invasive candidiasis remains substantial; after a decline in mortality throughout the early to mid 1990s, mortality rates have leveled off in recent years (Pfaller and Diekema 2007). There are a large number of well-characterized risk factors for invasive candidiasis, including (1) exposure to broad-spectrum antibiotics (Pfaller and Diekema 2007), (2) duration and degree of chemotherapy (Karabinis et al. 1988), (3) mucosal colonization (Pfaller and Diekema 2007), (4) indwelling vascular catheter (Diekema and Pfaller 2004), (5) total parenteral nutrition and severity of illness (Ostrosky-Zeichner 2003), (6) neutropenia (Prentice et al. 2000), (7) prior surgery (especially gastrointestinal) (Blumberg et al. 2001), (8) renal failure or 376 L. Putignani et al. Table 17.3 Main DNA-based assays for identification of medically important fungi other than Candida, Aspergillus and Cryptococcus spp. Target DNA Specimen Method Detection method Specificity References 18S rRNA Tissues and blood PCR Sequencing H. capsulatum Bialek et al. (2001) B. dermatitidis Bialek et al. (2003) 18S rRNA Tissue PCR Sequencing Rhizomucor spp. Bialek et al. (2005a, b) 18S rRNA Tissues PCR Sequencing 18S rRNA Culture PCR Ethidium bromide staining Trichosporon spp. Sugita et al. (1998) Cunninghamella spp. Kasai et al. (2008) 28S rRNA Plasma, respiratory (BAL) Real Time FRET and tissues Fusarium spp. Mishra et al. (2003) ITS Culture PCR Fuorescence Rhizopus spp. Nagao et al. (2005) ITS Tissues and serum PCR Ethidium bromide staining Fonsecaea pedrosoi Miyagi et al. (2008) ITS Culture PCR Sequencing ITS Culture Real Time FRET H. capsulatum Martagon-Villamil et al. (2003) Trichosporon spp. Sugita et al. (1999) ITS Culture PCR Sequencing Trichosporon asahii Sugita et al. (2001) ITS Serum PCR Ethidium bromide staining ITS2 Respiratory and tissue Real Time FRET Coccidioides spp. Binnicker et al. (2007) Trichosporon asahii Mekha et al. (2007) IGS1 Serum Real Time TaqMan Trichosporon spp. Rodriguez-Tudela et al. (2005) IGS1 Culture PCR Sequencing H. capsulatum Guedes et al. (2003) M antigen Culture PCR Ethidium bromide staining H. capsulatum Bracca et al. (2003) H antigen Blood and tissues PCR Ethidium bromide staining Hc100 Tissues PCR Ethidium bromide staining H. capsulatum Bialek et al. (2002a) Coccidioides posadasii Bialek et al. (2004) Ag2/PRA Tissues Real Time FRET Coccidioides posadasii Bialek et al. 2004 Ag2/PRA Tissues PCR Sequencing WI-1 Tissues PCR Sequencing B. dermatitidis Bialek et al. 2003 17 DNA-Based Detection of Human Pathogenic Fungi 377 hemodialysis (Blumberg et al. 2001), (9) bone marrow and solid-organ transplantation (Rüping et al. 2008), (10) recurrent or persistent gastrointestinal perforation (Eggimann et al. 1999), and (11) preterm delivery depending on gestational age and after birth intubation practices (Saiman et al. 2000; Wang et al. 2008). 17.6.1 Laboratory Diagnosis The challenging diagnosis of Candida infections reflects the complexity of the diseases sustained by these species. Traditional microbiological techniques for diagnosis of invasive candidiasis often fail to detect Candida spp. as blood cultures are often negative or become positive too late. Efforts to develop reliable diagnostic tests have stimulated the development of several serological methods for the diagnosis of Candida infection. However, antibody detection in patients with candidiasis is of limited usefulness for three reasons: (1) colonization by Candida spp. of the gastrointestinal tract or other sites can elicit antibody responses in uninfected individuals; (2) immunocompromised patients may not mount detectable antibody responses even when they have deep Candida infections; (3) development of a significant antibody titer occurs too late during infection. Morphologically, Candida spp. are thin walled and ovoid unicellular organisms (blastospores). Budding yeasts and pseudohyphae appear as Gram-positive. Candida organisms form smooth, creamy white, glistening colonies. Preliminary identification relies on Germ tube test, despite the fact that both false-positive and false-negative germ formation may occur (Sheppard et al. 2008). Traditional identification procedures are based on metabolic tests rather than on morphological characteristics (Wadlin et al. 1999). Because of variation in pathogenicity of individual species, accurate identification at the species level is highly recommended in clinical practice (Bishop et al. 2008; Sivakumar et al. 2009). In blood cultures, the growth kinetics may differ between Candida spp.: 1–3 days are necessary for C. albicans, C. parapsilosis, and C. tropicalis in standard medium, while growth of C. krusei and C. glabrata may take longer (Prevost and Bannister 1981). While candidemia would seem to be a key element of invasive candidiasis, retrospective studies have shown that blood cultures are positive in less than 50% of patients with autopsy-proven invasive candidiasis. (Rodriguez et al. 1996). Candiduria is common, especially in hospitalized patients having urinary catheters (Chen et al. 2008). However, neither the absolute colony count nor the presence or absence of white blood cells is pathognomonic (Navarro et al. 1997). It is particularly frequent to find Candida spp. in the respiratory tract of severely ill patients, but clinically relevant candidal pneumonia is quite rare and the presence of Candida spp. in the sputum has only a loose association with pneumonia (Rodriguez et al. 2000). Detection of Candida in CSF should always be regarded as pathognomonic of central nervous system (CNS) candidiasis. Despite intense and long-lasting efforts, no serological test has yet been shown to have clinical applicability. Very recently, a test employing Candida cytoplasmic 378 L. Putignani et al. antigens has been developed to measure circulating IgG, IgM, and IgA antibodies against C. albicans (Prince et al. 2008). Unlike the conventional antibody detection tests, the direct detection of Candida spp. antigens has been shown to have potential as an early diagnostic test. The Cand-Tec latex agglutination test (Ramco Laboratories,) was used as the first commercially available antigen detection test (Lemieux et al. 1990). However, the specificity and sensitivity of the Cand-Tec assay varied among reports (Phillips et al. 1990), and this test per se cannot establish a diagnosis of candidiasis. An alternative strategy is the detection of circulating b-(1-3)-Dglucan, a main cell wall component of Candida. High concentrations of b-(1-3)-Dglucan have been detected in patients with invasive candidiasis (Miyazaki et al. 1995a, b), and a commercial test is available (Fungitec G-test; Seikagaku Corporation). The detection of mannan antigenemia (mannanemia) for the immunodiagnosis of systemic candidiasis is widely used in patients with candidiasis, since positive mannan result may correlate with invasive candidiasis (Yeo and Wong 2002). Furthermore, correlation has been demonstrated between mannanemia and tissue invasion by Candida spp. in patients with candidemia, while mannanemia was less likely to be positive in patients with transient or central venous catheter-related candidemia (Girmenia et al. 1997). Two assays employing the anti-mamman EBCA1 monoclonal antibody are marketed as the Pastorex Candida latex agglutination test (Bio-Rad) and the Platelia Candida Antigen test (a double-sandwich enzyme immunoassay by Bio-Rad) (Sendid et al. 1999). Although the specificities of these two assays are similar, the EIA is more sensitive than the latex agglutination test (Sendid et al. 1999). Despite the attempt to improve the immunodiagnostic detection of mannan, most assays, like the Pastorex latex agglutination test, still lack sensitivity due to the rapid serum clearance of the antigen. 17.6.2 Molecular Detection The increasing incidence of Candida infections in immunocompromised patients (Cornely 2008) has focused attention on the exploitation of nucleic acids based techniques to set rapid and accurate diagnosis of IFI, independently from immunological-related markers. Nucleic acid hybridization and amplification methods provide both high detection rates and identification of Candida at the species level. This is increasingly important with the widespread use of antifungal therapy, and the problem of species-dependent resistance to antifungal agents in the genus Candida (Pfaller and Diekema 2004). Targets that are used in molecular diagnostic tests for Candida infections include both single and multicopy genes of nuclear and mitochondrial origin (Table 17.4). Among single copy genes of nuclear origin, actin, chitin synthase, cytochrome P450, and cytochrome P-450 lanosterol14a-demethylase (L1A1) have been exploited to detect Candida spp. by standard and nested-PCR from a wide set of clinical specimens including blood, serum, BAL, and body fluids (Kan 1993; Burgener-Kairuz et al. 1994; Chryssanthou et al. 1994; Jordan 1994) (Table 17.4). Species-specific restriction fragment length Method PCR Chitin synthase cytochrome P450 Blood Serum PCR PCR cytochrome P-450 lanosterol-14ademethylase (L1A1) L1A1 Blood, deep pus, peritoneal fluid, PCR pleural fluid, CSF, bile, urine, BAL BAL, blood PCR-RFLP L1A1 Blood PCR-RFLP 18S rRNA Culture PCR-RFLP HSP90 Swabs, urines, peritoneal fluid, pus, blood, serum Blood PCR-RFLP ITS 18S rRNA ITS Serum Serum Serum PCR-EIA PCR PCR 25–28S rRNA LSU D2/D1 ITS1-5.8S-ITS2, 25–28S rRNA LSU D2 Repetitive sequences 25–28S rRNA LSU D2 Culture EO3,duplicated mitochondrial region Detection method Hybridization with radiolabeled probe Southern blotting Ethidium bromide staining Southern blotting Specificity Candida spp. References Kan (1993) Candida spp. Candida albicans Jordan (1994) Chryssanthou et al. (1994) Candida spp. Burgener-Kairuz et al. (1994) Candida spp. Morace et al. (1997) Candida spp. Morace et al. (1999) Candida spp. Hopfer et al. (1993) C. albicans Crampin and Matthews (1993) C. albicans Miyakawa et al. (1993) Candida spp. Candida spp. Candida spp. Burnie et al. (1997) Einsele et al. (1997) Bougnoux et al. (1999) PCR Ethidium bromide staining/Southern blotting Ethidium bromide staining Ethidium bromide staining Hybridization with radiolabeled probe Ethidium bromide staining/Southern blotting Enzyme immunoassay Southern blotting Ethidium bromide staining Sequencing Candida spp. Linton et al. (2007) Culture PCR Sequencing Candida spp. Sanguinetti et al. (2007) Culture Culture Rep-PCR PCR Microfluidics chip Sequencing Candida spp. Wise et al. (2007) Hall et al. (2003) PCR (continued) 379 Specimen Serum DNA-Based Detection of Human Pathogenic Fungi Target DNA Actin 17 Table 17.4 Main DNA-based assays for identification of Candida spp Specimen Culture Method PCR Detection method Sequencing Blood culture PCR ITS1 and ITS2 Blood culture Multiplex-PCR EF3, CDC3, HIS3 microsatellite M13 minisatellite Culture Real-Time Ethidium bromide Candida spp. staining Candida spp. Ethidium bromide staining Primer fluoro-labeling C. albicans Culture PCR ITS2 25-28S rRNA LSU D2/D1 ITS Culture Culture PCR LAMP Blood culture Real-Time Septifast ITS2 RNA subunit of RNase P ITS2 ITS2 18S rRNA Blood culture Blood Blood Culture Blood culture Real-Time Real-Time Real-Time Real-Time Real-Time Melting curve TaqMan TaqMan TaqMan TaqMan ITS1 and ITS2 18S rRNA gene Serum Culture Real-Time Real-Time Melting curve Melting curve Ethidium bromide staining Pyrosequencing LAMP amplicon DIGlabeling Melting curve Specificity Candida spp. References Putignani et al. (2008a, b) Li et al. (2003) Chang et al. (2001) Beretta et al. (2006) C. albicans Bartie et al. (2001) Candida spp. Candida spp. Boyanton et al. (2008) Inácio et al. (2008) C. albicans, C. tropicalis, C. parapsilosis, C. krusei, C. glabrata Candida spp. Candida spp. Candida spp. Candida spp. FLC-sensitive species (C. albicans, C. tropicalis, C. parapsilosis, C. dubliniensis) and FLC-resistant (C. glabrata, C. krusei) Candida spp. Candida spp. Mancini et al. (2008) Selvarangan et al. (2003) Innings et al. (2007) Maaroufi et al. (2003) Guiver et al. (2001) Metwally et al. (2007) Dunyach et al. (2008) White et al. (2004) L. Putignani et al. Target DNA 25–28S rRNA LSU D2 ITS1 and ITS2 380 Table 17.4 (continued) 17 DNA-Based Detection of Human Pathogenic Fungi 381 polymorphisms (RFLPs) have been identified in the L1A1 locus, making possible the identification of Candida species directly from DNA extracted from BAL and blood, even though traditional blood cultures and antigen detection assays were negative (Morace et al. 1997, 1999). Also HSP90 and ribosomal DNA RFLPs patterns have been exploited to perform identification of C. albicans and Candida spp. from clinical specimens (Crampin and Matthews 1993; Hopfer et al. 1993). Single-strand conformational polymorphisms of the 18S rDNA have provided differential profiles for Candida spp. and several yeast and mold species (Walsh et al. 1995). Methods targeting multicopy genes offer lower detection limit in terms of number of fungal genomes. Among multicopy genes, mitochondrial DNA has been used in the PCR-based detection of C. albicans (Miyakawa et al. 1992) and Candida spp. (Yokoyama et al. 2000). However, the variability of mitochondrial DNA among different strains may be a limiting factor. Other studies have targeted the multicopy rDNA gene cluster with universal primer sets, in order to maximize sensitivity and specificity (Sandhu et al. 1995; Martin et al. 2000, Putignani et al. 2008a, b). As already discussed, the ribosomal genes contain conserved sequences that are common to all fungi (Figure 17.1), and which can be used to screen for yeast presence (Burnie et al. 1997; Einsele et al. 1997), while the variable ITS and LSU D2 sequences can be exploited for species identification (Linton et al. 2007; Sanguinetti et al. 2007; Putignani et al. 2008a, b). Indeed, the commercial MicroSeq D2 LSU rDNA Fungal Identification Kit can easily be applied to yeast identification (Hall et al. 2003). Since non-albicans Candida spp. are increasing in importance (Bille et al. 2005; Tavanti et al. 2005), broad-range diagnostic approaches capable of identifying a large number of Candida species are required (Table 17.4). The huge amount of sequencing data generated for the ribosomal 25–28S rDNA target have recently allowed molecular mycologists to compare between DNA-based identification procedures and classical diagnostic methods using large cohorts of patients and yeast isolates (Linton et al. 2007; Sanguinetti et al. 2007). New technologies, such as the pyrosequencing-based method, have recently been developed to perform identification of Candida spp. (Boyanton et al. 2008). Moreover, an innovative technique alternative to the PCR and based on isothermal DNA amplification, is providing highly performing identification of Candida (Inácio et al. 2008). Repeated genome regions, dispersed through the fungal genomes, have also been considered to assess inter and intraspecies variability within the genus Candida. A real-time PCR study performed by using EF3, CDC3, and HIS3 microsatellite sequences provided an interesting example of genotyping clustering for cases of candidaemia in an intensive care unit (Beretta et al. 2006). Also a minisatellitespecific M13 primer was exploited to assess a mixed population of C. albicans strains in the oral microflora of patients affected by chronic hyperplastic candidosis (CHC) (Bartie et al. 2001). More recently, the DiversiLab rep-PCR-based system made it possible to identify and differentiate clinical isolates of Candida spp. (Wise et al. 2007). Most of the questions raised on the value of PCR assays for invasive aspergillosis apply also to disseminated candidosis (Bretagne and Costa 2005). The aim of several reports using PCR assays (Table 17.4) is to demonstrate the 382 L. Putignani et al. superiority of PCR assays over blood cultures. The latter are known to be poorly (ca. 50%) sensitive, as previously discussed. A likely explanation is the low burden of circulating yeasts makes unlikely their sampling from blood. Therefore, the amplification of Candida DNA from blood raises the question whether the DNA comes from living Candida or it is naked DNA from lysed cells. If the aim of PCR assays is to improve the detection of viable Candida cells, this should be related to the sensitivity of blood culture, usually comprised between 1 and 150 CFU/mL (Einsele et al. 1997). With currently used DNA extraction kits, the input blood volume is usually 200 mL. Thus, in order to obtain a PCR-positive result, at least ten Candida cells must be present in the 200 mL tested (i.e., 50 cells/mL), since only 1/10th of the extracted material is loaded in the PCR reaction. Under these conditions, a blood culture performed with 10 mL of blood is expected to turn positive unless the Candida fails to grow because of antifungal treatment. If the amplified DNA comes from circulating nacked DNA, as suggested by the better yield reported for serum than for blood (Burnie et al. 1997; Bougnoux et al. 1999), the sensitivity using multicopy genes as a target could reach higher sensitivity than 50 Candida cells/mL, probably below one Candida genome/mL assuming that the target is > 50 copies per cell, as in case of the rDNA. Indeed, the sensitivity of the real-time PCR (Light Cycler) system targeting the ITS was estimated ca. 1 cell/mL (White et al. 2003). However, the meaning of such a finding is less clear than a positive blood culture, especially in heavily colonized patients, as seen in ICUs, or patient under antifungal therapy. Another potential limit is the need to detect several Candida species, at least the five or six main species encountered in blood cultures (Tortorano et al. 2004). Thus, a good method should distinguish between most of the Candida spp. which are sensitive to fluconazole, and C. glabrata and C. krusei whose sensitivity is variable or null. Detection of anonymous yeasts in a pathological sample, without the knowledge of the involved species, cannot direct antifungal therapy until yeast identification. Therefore, to overpass the traditional culture on agar media, a realtime PCR assay should enable to detect and identify every Candida spp. even in polymicrobial association. The generation of new real-time PCR instruments which can simultaneously use up to four fluorogenic probes (as in the Light Cycler SeptiFast system) is a substantial advance in this direction. Another option is to associate PCR assays with a DNA-chip technology as in the rep-PCR DiversiLab system. As for Aspergillus, the use of real-time PCR should improve the reproducibility of the PCR tests and make comparison of the results from several studies feasible. The current assays are methodologically heterogeneous (Table 17.4). Recently, panfungal-primer based assays (Table 17.1) have also been applied to solve emerging clinical issues, as the detection of invasive candidosis in renal transplant recipients (Badiee et al. 2007) (Table 17.1). 17.7 Coccidioides spp. Coccidioides immitis and Coccidiodes posadasii, the only species included in the genus Coccidioides, are dimorphic fungi found in soil particularly in warm and dry areas with low rain fall, high summer temperatures, and low altitude. The two 17 DNA-Based Detection of Human Pathogenic Fungi 383 species are morphologically identical but genetically and epidemiologically distinct (Fisher et al. 2001, 2002). C. immitis is geographically limited to California’s San Joaquin valley region, whereas C. posadasii is found in the southwest of the United States, Mexico, and South America. Imported cases are observed following travel to endemic areas (Cairns et al. 2000). The two species co-exist in the southwest of the United States and Mexico. Coccidioides spp. are causative agent of coccidioidomycosis in humans. Coccidioidomycosis is a true systemic mycoses (Galgiani 1999), acquired by inhalation, and initially presents with a pulmonary infection which may later disseminate to other organs and systems. Airway coccidioidomycosis involving the endotracheal and endobronchial tissues may develop (Polesky et al. 1999). The clinical picture has a remarkably wide spectrum. The infection remains as an acute and self-limiting respiratory infection in most exposed individuals, but it can progress to a chronic and sometimes fatal disease in others. Spontaneous healing is observed in ca. 95% of the otherwise healthy individuals. Clinical presentations of coccidioidal infection are acute pneumonia, chronic progressive pneumonia, pulmonary nodules and cavities, extrapulmonary nonmeningeal disease, and meningitis (Chiller et al. 2003). Due to the true pathogenic nature of the fungus, coccidioidomycosis affects otherwise healthy, immunocompetent humans, although it may also affect immunocompromised patients, such as AIDS patients and organ transplant recipients (Medoff et al. 1992; Blair and Logan 2001). In fact, concurrent risk factors are HIV infection, organ transplant, hematologic malignancy and pregnancy (Powell et al. 1983). Coccidioidomycosis is a common cause of community-acquired pneumonia (CAP) in disease-endemic areas. However, because Coccidioides spp. testing among CAP patients is infrequent, reportable-disease data, which rely on positive diagnostic test results, greatly underestimate the true disease prevalence (Chang et al. 2008). 17.7.1 Laboratory Diagnosis The diagnosis of coccidioidal infection can be traditionally made in three ways: (1) identification of coccidioidal spherules in a cytology or biopsy specimen, (2) culture from any body fluid that is positive for Coccidioides spp., or (3) a serologic test that is positive for the Coccidioides spp. Since Coccidioides does not colonize humans, the finding of spherules in tissue, sputum, bronchoalveolar lavage fluid, or other body fluid or a positive culture from any location in the body is pathognomonic of coccidioidal infection. For the safe isolation of Coccidioides spp., the laboratory should maintain a biological safety level 2 or 3. Serodiagnosis can be used to detect coccidioidal infection. Early immune response is characterized by the presence of IgM, which can be detected by a tube precipitin method, immunodiffusion, latex agglutination, or enzyme immunoassay (EIA). Latex agglutination and EIA are highly sensitive but are associated with false-positive results (Pappagianis 2001). These qualitative tests provide positive or negative results but no quantitative information. By comparison, complement fixation provides a quantitative titer that reflects the intensity of the immune response (Galgiani 1992). 384 L. Putignani et al. 17.7.2 Molecular Detection Identification of coccidioidal elements in tissue sections can be very difficult or impossible (Kaufman et al. 1998). A successful approach to C. immitis detection in paraffin-embedded tissue sections is in situ hybridization. Hayden et al. (Hayden et al. 2001) described a set of oligonucleotide probes which identify and differentiate yeast like organisms in tissue sections. The C. immitis-specific probe had a sensitivity of 94.3%, a 100% specificity, and a positive predictive value of 100%. However, identification by this method is limited to cases with microscopically visible fungal elements. Stockman et al. (1993) showed a commercially available acridinium ester-labeled chemiluminescent DNA probe targeting the ribosomal RNA of Coccidioides spp. to be sensitive and 100% specific. A total of 164 strains from related and unrelated fungal species were tested to define specificity, and no cross-reaction was detected. The probe was developed for detection on spherule in tissue samples. Fixation in formaldehyde reduces efficiency of this excellent and widely used identification system (Gromadzki and Chaturvedi 2000). Nested PCR and a real-time PCR assays were recently developed to target the genus-specific antigen2/proline rich antigen of Coccidioides spp. (Bialek et al. 2004) (Table 17.3). Melting curve analysis by LightCycler and sequencing of the 526-bp product of the first PCR correctly identified all strains as C. posadasii. In addition, specific DNA was amplified by the conventional nested PCR from three microscopically spherulepositive paraffin-embedded tissue samples whereas 20 human samples positive for other dimorphic fungi remained negative (Bialek et al. 2004). Another LightCycler assay, targeting the ITS2 locus, allows identification of Coccidioides spp. from various respiratory specimens (Binnicker et al. 2007). Finally, a rep-PCR assay, binding to multiple noncoding, repetitive sequences interspersed throughout the genome, was exploited to test different Coccidiodes spp. (Pounder et al. 2006) (Table 17.3). Distinction between C. inmitis and C. posadasii species was assessed by designing appropriate coupled primers amplifying nucleotides 660.313–661.032 of C. immitis contig 2.2 (AAEC02000002, (http://www.broad.mit.edu/annotation/ fungi/coccidioides_immitis/). This test discriminates C. posadasii from C. immitis for a 86-bp deletion (Umeyama et al. 2006). 17.8 Cryptococcus neoformans Although more than 30 species are included in the genus Cryptococcus, only two of them are pathogenic: Cryptococcus neoformans and Cryptococcus gattii (KwonChung et al. 2002). In the environment, C. neoformans is primarily found associated with fecal excretions from certain birds, such as pigeons, and in tree hollows. For years, C. gattii was found primarily in tropical and subtropical regions. It has been associated primarily with eucalyptus trees, which were considered its primary niche (Hull and Heitman 2002). These species were previously classified as three 17 DNA-Based Detection of Human Pathogenic Fungi 385 varieties: C. neoformans var neoformans, C. neoformans var grubii, and C. neoformans var gattii (Franzot et al. 1999), which were classified into five capsular serotypes: A, D, and the hybrid diploid AD belonging to the C. neoformans, with serotype A named C. neoformans var grubii, and serotype D named C. neoformans var neoformans; and serotype B and C classified as C. gattii (Franzot et al. 1999) and eight molecular genotypes (VNI through VNIV for C. neoformans and VGI through VGIV for C. gattii) (Meyer et al. 2003). Principal predisposing factors for cryptococcosis are HIV infection, treatment with corticosteroids, solid organ transplantation with immunosuppressive therapies, malignancies, CD4+ T-cell lymphopenia, connective tissue diseases or immunologic diseases, diabetes mellitus, chronic pulmonary diseases or lung cancer, renal failure or peritoneal dialysis, cirrhosis and pregnancy (Perfect and Casadevall 2002). Before the HIV epidemic, cryptococcal infection was an uncommon systemic fungal infection that occurred primarily in patients who had impaired immunity (Mitchell and Perfect 1995). However, during the past two decades of the HIV epidemic, the incidence of cryptococcosis increased dramatically. Before the development of the “highly active antiretroviral therapy” (HAART), cryptococcal infection was regarded as the major cause for morbidity and mortality in HIV-infected patients with CD4 lymphocyte counts < 100 cells/mL. However, in the HAART era, the incidence of cryptococcosis decreased significantly in HIV patients, although remaining constant in non-HIV individuals (Friedman et al. 2005). A recent clinical syndrome associated with HAART-driven immune reconstitution in HIV patients is “the immune reconstitution inflammatory syndrome” (IRIS). IRIS has been reported to occur in 30–35% of HIV patients with a history of cryptococcosis in whom HAART was initiated (Perfect and Casadevall 2002). Patients infected with HIV who have cryptococcal meningitis and IRIS have a greater fungal burden, as indicated by higher CSF antigen titer and the presence of disseminated infection or fungemia. Nevertheless, cryptococci are hardly culturable from these patients, and molecular diagnosis is recommended (Putignani et al. 2008a, b). Similar to patients infected with HIV, those who have undergone solid organ transplantation and having higher cryptococcal antigen titers with disseminated disease are more likely to develop IRIS after initiation of antifungal therapy (Singh et al. 2005). Cryptococcosis is the third most common invasive fungal infection after candidiasis and aspergillosis in patients undergoing solid organ transplantation (Vilchez et al. 2002). The CNS and respiratory tract are the most plagued organs by C. neoformans and C. gattii infections. 17.8.1 Laboratory Diagnosis Traditionally, diagnosis of cryptococcosis depends on Cryptococcus culture or on the demonstration of encapsulated yeasts in India ink-stained pathological samples. Direct microscopy and culture are specific but the sensitivity is poor (50–80%) (Snow et al. 1975). Although culture remains the “gold standard” for diagnosis, it 386 L. Putignani et al. is cumbersome, labor intensive, and time consuming because of the slow growth of cryptococci. On the other hand, negative cultures may occur despite positive India ink examinations because of nonviable yeast cells that may have prolonged persistence at the infection site (Putignani et al. 2008a, b). Serology is an indirect and adjunct or complementary procedure to support clinical diagnosis, especially when the patient is on treatment. Antigen detection still represents the fastest and simplest diagnostic tool. Detection of cryptococcal capsular polysaccharide antigen in serum or body fluids has performed robustly for many years. The main component of the C. neoformans capsular polysaccharide is a glucuronoxylomannan (GXM). Antigenic structures intrinsic to the GXM allow the distinction between serotypes. An important tool in the diagnosis of cryptococcosis is the latex particle agglutination test (LAT, e.g., Cryptococcus-Antigen Latex-Agglutination System; CALAS, Meridian Bioscience Inc.), which uses latex particles coated with an anti-GXM antibody to detect capsular polysaccharide antigen in serum or CSF. LAT is the most commonly used serological method due to its simplicity in performance (Hamilton et al. 1991; Kiska et al. 1994), although suffering from false positivity and difficulty of interpretation in borderline cases (Whittier et al. 1994; Millon et al. 1995). Enzyme immunoassay (EIA, e.g., PREMIER Cryptococcal Antigen Kit; Meridian Bioscience) is another serological tool for detection of capsular polysaccharide antigens of C. neoformans in CSF. This is a rapid test that provides visual and numeric result in less than an hour without pre-treatment of the specimen (Saha et al. 2008). 17.8.2 Molecular Detection Various DNA extraction procedures have been published for efficient disruption of cryptococcal cells, including enzyme digestion or glass beads (Tanaka et al. 1996), and a number of PCR assays have successfully been applied to diagnose cryptococcal disease starting from a variety of clinical specimens, such as blood, liquor, secretions, cutaneous scrapings, bronchial alveolar aspirate, and urine. These PCR assays use primers targeting the 18S, 28S, or the ITS and 5.8S rDNA, eventually coupled with species-specific probes (Mitchell et al. 1994; Sandhu et al. 1995; Prariyachatigul et al. 1996; Rappelli et al. 1998; Kano et al. 2001; Bialek et al. 2002b; Iyer and Banker 2002; Takahashi et al. 2003; Pagano et al. 2004; Paschoal et al. 2004; Putignani et al. 2008a, b) (Table 17.5). Unique primers named CN4, CN5, and CN6, have been developed to amplify the 5.8S and ITS regions of C. neoformans rDNA, and used with different clinical specimens (Mitchell et al. 1994). Rappelli et al. (1998) set up a nested PCR assay on the ITS region of C. neoformans directly in CSF specimens, followed by visual detection in agarose gels (Rappelli et al. 1998). This method was improved by Paschoal et al. (2004), who shortened the time for C. neoformans detection and identification from CSF samples by entailing just a single step for the amplification. The results demonstrated that PCR had the highest sensitivity rate (92.9%), superior to culture (85.7%) and to India ink test (76.8%) (Paschoal et al. 2004). Rapid cycling 17 DNA-Based Detection of Human Pathogenic Fungi 387 Table 17.5 Main DNA-based assays for identification of C. neoformans Target DNA Specimen Method Detection method References 18S rRNA Tissues PCR Sequencing Bialek et al. (2002b) 18S rRNA Tissues Real Time Real time (FRET) Bialek et al. (2002b) 18S rRNA CSF PCR Southern Prariyachatigul et al. (1996) hybridization 18S rRNA Culture PCR Ethidium bromide Mitchell et al. (1994) staining ITS2 CSF PCR Ethidium bromide Rappelli et al. (1998) staining ITS2 Lymph node PCR Ethidium bromide Putignani et al. (2008a, b) aspirate staining ITS2 Respiratory PCR Sequencing Takahashi et al. (2003) (BAL) Culture PCR Sequencing Putignani et al. (2008a, b) 25–28S rRNA LSU D2 ITS5.8S CSF PCR Ethidium bromide Paschoal et al. (2004) rRNA staining CAP59 Tissues PCR Ethidium bromide Kano et al. (2001) staining Ethidium bromide Tanaka et al. (1996) URA5 Respiratory PCR and staining tissues real-time PCR protocols further simplified the diagnostic laboratory workflow and reduced the possibility of product contaminations (Bialek et al. 2002b; Bergman et al. 2007). Bialek et al. (2002b) established two PCR protocols targeting the 18S rRNA gene of C. neoformans. One protocol was designed as a nested PCR to be performed in conventional thermal cyclers. However, to minimize the event of false-positive results, amplicons should be sequenced for an unambiguous species identification. The other protocol was designed as a quantitative single-round PCR adapted to LightCycler technology (Bialek et al. 2002b) which avoids amplicon verification by means of Southern blotting or DNA sequencing. Once DNA is extracted from suitable specimens and reaction mixtures are completed, the results of sensitive and quantitative PCR are available within 60 min (Bialek et al. 2002b). A simple PCR-based method for C. neoformans serotyping strains, which uses a set of four primers for the laccase (LAC1) gene has recently been developed (Ito-Kuwa et al. 2007). This primer combination differentiates serotypes A, D, B, and C. Further differentiation between serotypes AD and D requires the use of a primer pair to the capsule (CAP64) gene. When multiplex PCR is performed with all of the above six primers, the five serotypes generate distinct fingerprints composed of two to five fragments. Genotype-based differentiation of the C. neoformans serotypes can further be achieved by combining PCR-RFLP analysis of the CAP10 and CAP59 genes (Raimondi et al. 2007). A reverse line blot hybridization panfungal assay for identification of C. neoformans isolates is also available (Zeng et al. 2007) (Table 17.1). 388 17.9 L. Putignani et al. Dematiaceous Fungi Chromoblastomycosis, mycetoma, and phaeohyphomycosis are fungal infections caused by dematiaceous (darkly-pigmented) fungi, a group of organism usually found in the soil (Revankar et al. 2002). Chromoblastomycosis, a chronic infection of skin and subcutaneous tissues, is most commonly seen in tropical areas with most cases caused by Fonsecaea pedrosoi followed by Fonsecaea compacta, Phialophora verrucosa, Cladophialophora carrionii, and Rhinocladiella aquaspersa (Revankar et al. 2002, Sanche et al. 2003). Infection typically occurs via traumatic implantation in exposed surfaces of the legs (Milam and Fenske 1989). Mycetoma, also known as Madura’s foot, is also a chronic infection of cutaneous and subcutaneous tissues, caused by Madurella mycetomatis, Madurella grisea, Curvularia lunata, Exophiala jeanselmei, and Leptosphaeria senegalensis (Sanche et al. 2003). Chromoblastomycosis and mycetoma are considered types of phaeohyphomycoses, which can also include corneal, systemic infections, and fulminant disseminated disease (Revankar et al. 2002; Sanche et al. 2003). Over 100 species have been implicated in phaeohyphomycoses (Revankar et al. 2002). Genera associated with pneumonia include Ochroconis, Exophiala, and Chaetomium (Revankar et al. 2004) while Scedosporium (e.g., S. prolificans) and, to a lesser extent, Bipolaris and Wangiella produce disseminated disease (Revankar et al. 2002, 2004). Genera associated with CNS infection include Cladophialophora (e.g., C. bantiana), Ramichloridium (e.g., R. mackenzei), and Ochroconis (Kantarcioglu and de Hoog 2004; Revankar et al. 2004; Revankar 2006). In contrast to other IFIs, underlying immunodeficiency is not a prerequisite for phaeohyphomycoses, although some immune dysfunction is associated with disseminated disease (Revankar et al. 2002). Mortality rates are high regardless of the patient’s immune status; however, recovery from neutropenia is considered critical for patients with S. prolificans infection (Revankar et al. 2002). No risk factors have been identified in many patients (Kantarcioglu and de Hoog 2004; Revankar et al. 2004). However, CNS disease is generally correlated to cellular immune dysfunction while disseminated disease to malignancy, neutropenia, HSCT, solid organ transplant (SOT), HIV, and catheterism (Cornely 2008). 17.9.1 Laboratory Diagnosis Classical methods are based on 20% potassium hydroxide microscopy, histopathological confirmation of sclerotic cells by periodic acid-Schiff stain, culture on Sabouraud’s glucose agar, slide culture method, and observation of conidia by scanning electron microscopic examination (Hospenthal 1995). 17.9.2 Molecular Detection DNA-based techniques have recently been employed to identify the causative agents of chromoblastomycosis (Vidal et al. 2004; Piepenbring et al. 2007; 17 DNA-Based Detection of Human Pathogenic Fungi 389 Chowdhary et al. 2008; Miyagi et al. 2008). PCR of the ITS and direct sequence analysis of the amplicon, coupled with classical detection methods proved effective in identification of the causative agent of chromoblastomycosis as Fonsecaea pedrosoi (Miyagi et al. 2008). By means of 18S rDNA sequencing and GenBank database homology searches, a rare case of chromoblastomycosis in a renal transplant recipient caused by a nonsporulating fungal isolate belonging to genus Rhytidhysteron was successfully diagnosed (Chowdhary et al. 2008). 17.10 Fusarium spp. Fusariosis is a life-threatening mycosis in immunocompromised hosts (Lionakis and Kontoyiannis 2004). Fusarium spp. are angiotropic and angioinvasive molds that produce hemorrhagic infarction and low tissue perfusion, resulting in tissue necrosis (Lionakis and Kontoyiannis 2004). More than 50 species of Fusarium have been identified but only a few are pathogenic in humans (Lionakis and Kontoyiannis 2004). These include F. solani (causes ~ 50% of cases), Fusarium oxysporum, Fusarium moniliforme, Fusarium verticillioides, Fusarium dimerum, and Fusarium proliferatum (Lionakis and Kontoyiannis 2004). In terms of global occurrence, fusariosis is most common in the United States (50–80% of all fusariosis cases), followed by France, Italy, and Brazil (Torres et al. 2003). Invasive fusariosis has emerged in many tertiary-care cancer centers as the second most common invasive mold infection (after invasive aspergillosis) in severely immunocompromised patients (Lionakis and Kontoyiannis 2004). A steady rise in the number of cases of fusariosis in hematopoietic stem cell transplant recipients has been observed since the late 1980s (Nucci et al. 2004). More than 90% of cases of fusariosis have been reported in neutropenic patients with hematologic malignancies (Lionakis and Kontoyiannis 2004) and autologous bone marrow transplant recipients (Boutati and Anaissie 1997). Because the clinical presentation of fusariosis is rather unspecific, differentiation from invasive aspergillosis can be challenging (Torres et al. 2003; Lionakis and Kontoyiannis 2004). 17.10.1 Laboratory Diagnosis Fusarium spp. grow easily and rapidly in most mycological media. Although the genus Fusarium can be identified by the production of hyaline, banana-shaped, multicellular macroconidia with a foot cell at the base, interpretation of the growth of Fusarium spp. from different biological materials depends on the clinical context. The clinician and the microbiologist must be cautious, because Fusarium spp. may contaminate laboratory specimens and pseudo-outbreaks of fusariosis may occur (Grigis et al. 2000). In support of infection are the isolation of several colonies from the same specimen or of the same fungus from different specimens of the same patient, the site of isolation, and, most importantly, a positive direct examination of the biological material. Histopathology is therefore recommended 390 L. Putignani et al. for a confirmatory diagnosis of fusariosis. In tissues, the hyphae of Fusarium are similar to those of Aspergillus spp. (hyaline and septate filaments). However, adventitious sporulation may be present in tissue, and the finding of both hyphae and yeast-like structures is highly suggestive of fusariosis in the high-risk population. In the absence of microbial growth, distinguishing fusariosis from other hyalohyphomycoses may be difficult and requires the use of molecular techniques. 17.10.2 Molecular Detection A fluorescent-based PCR assay which allows rapid and reliable identification of five toxigenic and pathogenic Fusarium spp., namely Fusarium avenaceum, F. culmorum, F. equiseti, F. Oxysporum, and F. sambucinum has been developed (Mishra et al. 2003). This method is based on amplification of species-specific ITS using fluorescent oligonucleotide primers, which were designed on sequence divergence within the rDNA. Besides providing an accurate, reliable, and rapid identification of these Fusarium spp., it reduces the potential for exposure to carcinogenic chemicals as it substitutes ethidium bromide staining of amplicons with the use of fluorescent dyes (Mishra et al. 2003). Molecular diagnosis of the toxigenic Fusarium spp. has recently been developed in a PCR assays for genes involved in the toxin biosynthesis (Mulè et al. 2005). The method allows early detection of toxin-producing Fusarium spp, and reveals which particular toxin may be present in a feed product (Mulè et al. 2005). A DNA microarray was assembled to detect and identify DNA from 14 fungal pathogens including F. oxysporum and F. solani in blood, BAL, and tissue samples from high-risk patients. The assay combines multiplex PCR and consecutive DNA microarray hybridization. PCR primers and capture probes were derived from unique sequences of the 18S, 5.8S, and ITS1 regions of the fungal rRNA genes. Hybridization with genomic DNA of fungal species resulted in species-specific hybridization patterns (Spiess et al. 2007). 17.11 Histoplasma capsulatum H. capsulatum is a dimorphic fungus that, although worldwide distributed, is more prevalent in certain regions of north and central America (Wheat and Kauffman 2003). Manifestations of histoplasmosis range from asymptomatic to severe and fatal disease, depending on the immune status and the magnitude of exposure (Wheat and Kauffman 2003). Disseminated disease occurs in approximately 10% of all infections (Wheat and Kauffman 2003). CNS involvement is the presenting manifestation in 5–10% of disseminated disease cases (Wheat and Kauffman 2003). Although histoplasmosis is relatively uncommon among transplant patients, outbreaks of the disease have been reported in transplant centers with incidence rates of ca. 2% (Freifeld et al. 2005). While immunocompetent adults with disseminated histoplasmosis typically have a chronic progressive course of the disease, those who are severely immunosuppressed can have an acute course with fatal outcome. 17 DNA-Based Detection of Human Pathogenic Fungi 17.11.1 391 Laboratory Diagnosis Since the clinical picture of histoplasmosis strongly mimics those of tuberculosis and some other lung diseases, it is not possible to confirm a diagnosis of pulmonary histoplasmosis on the sole basis of the clinical symptoms (Bracca et al. 2003). Confirmatory diagnosis requires culture, fungal stain of tissue or body fluids, and tests for antibodies and antigens, depending on the extent and severity of infection (Wheat 2003). Microscopic identification of Histoplasma in May-Grünwald-Giemsa-stained slides is an acceptable approach, but an experienced operator is needed to obtain reliable results. The fungus can be cultured from different clinical samples such as blood, bone marrow, respiratory secretions, or localized skin lesions (Wheat 2003). Isolation of the organism from cultures provides the strongest evidence for infection with H. capsulatum, but has some limitations. First, growth of H. capsulatum in culture is slow, requiring up to 4 weeks, causing undesirable delay in the diagnosis and therapy. Second, invasive procedures might be required to obtain specimens for culture. Bone marrow cultures yield the highest frequency of positive isolations (70–90%), while respiratory specimens yield positive results less frequently (50–90%). Third, cultures are negative in most patients with mild forms of histoplasmosis, limiting its use for diagnosis in mild cases. Cultures are positive primarily in patients with disseminated or chronic pulmonary histoplasmosis, but even in these patients cultures can be falsely negative in about 20% of disseminated histoplasmosis and 50% of chronic pulmonary diseases (Wheat 2003). Serological methods are faster than culture, but they can lead to false-positive results because the titer of specific antibodies against Histoplasma remains high for months or years after the primary infection (Bullok 1995). In addition, cross-reactivity against P. brasiliensis can give false positive results (Raman et al. 1990). On the other hand, false-negative results due to low antibody titers can be observed in immunocompromised patients with active infection (Wheat et al. 1992). An alternative is the detection of H. capsulatum antigen in body fluids (Gomez et al. 1997). Although this method is more sensitive than serum antibody detection assays, cross-reactivity against antigens of other pathogenic fungi remains a problem (Wheat et al. 2002). 17.11.2 Molecular Detection A variety of PCR and DNA probes have been applied to the detection of H. capsulatum DNA in tissues and body fluids (Sandhu et al. 1995; Bialek et al. 2005a; Rickerts et al. 2002; Bracca et al. 2003; Martagon-Villamil et al. 2003). A seminested PCR for the diagnosis of histoplasmosis that amplifies a portion of the H. capsulatum H antigen gene has been developed. This assay is sensitive and specific, being able to detect genomic material corresponding to less than ten yeast cells without cross-reaction against other bacterial or fungal pathogens (Bracca 392 L. Putignani et al. et al. 2003). The real-time PCR assay for H. capsulatum can be used for the confirmation of culture isolates suspected to be H. capsulatum, and potentially to test clinical specimens directly (Martagon-Villamil et al. 2003). A rep-PCR assay, targeting multiple noncoding, repetitive sequences (generally 30–500 bp) interspersed throughout the fungal genome, was exploited to differentiate isolates of H. capsulatum, according to the procedure previously described for B. dermatitidis (Pounder et al. 2006) (Table 17.3). 17.12 Trichosporon spp. Trichosporonosis is a relatively uncommon opportunistic fungal infection in immunocompromised individuals, often resulting in fatal outcome (Girmenia et al. 2005). The taxonomy of the yeasts that causes trichosporonosis has recently been revised. It is now widely accepted that the previously named T. beigelii consists actually of six species: Trichosporon asahii, Trichosporon asteroides, Trichosporon cutaneum, Trichosporon inkin, Trichosporon mucoides, and Trichosporon ovoides. Geotrichum capitatum, originally included in genus Trichosporon, has been classified as Blastoschizomyces capitatus, whose Dipodascus capitatus is the teleomorph stage. B. capitatus is also a common cause of trichosporonosis (Girmenia et al. 2005) and has recently been characterized as an outbreak agent (Ersoz et al. 2004). Case reports of invasive trichosporonosis have frequently been reported, and Trichosporon and Blastoschizomyces are currently considered the second most common yeasts causing disseminated infections after Candida spp. (Chagas-Neto et al. 2008). Trichosporonosis is an acute, febrile, severe infection with dissemination to multiple deep organs, associated with up to 64% mortality (Chagas-Neto et al. 2008). While any immunocompromised patient can develop trichosporonosis, the risk is highest for those with hematologic malignancies, followed by neutropenia, peritoneal dialysis, solid tumor, SOT, burns, and prosthetic cardiac valve (Girmenia et al. 2005). Incidence rates of 0.4% and 0.5%, respectively, for infections due to Trichosporon spp. and G. capitatum have been reported in patients with leukemia (Girmenia et al. 2005). Trichosporon spp. are resistant to echinocandins, as demonstrated by breakthrough trichosporonosis in immunocompromised patients treated with caspofungin or micafungin (Cornely 2008). Trichosporon infection should be suspected in immunocompromised patients who are being treated with echinocandins and develop signs of septicemia or local infection. 17.12.1 Laboratory Diagnosis A firm diagnosis of deep-seated trichosporonosis should result from histological examination of tissue samples obtained by biopsy and culture of living Trichosporon cells. However, invasive examinations, such as biopsy, cannot repeatedly be 17 DNA-Based Detection of Human Pathogenic Fungi 393 performed in patients with severe underlying diseases, and even though the detection of causative fungi in the blood is clinically valuable, the results of blood cultures usually become available after a long time. Patients with disseminated trichosporonosis may test positive for the GXM antigen (Melcher et al. 1991) or (1 ! 3)-b-D-glucan (Miyazaki et al. 1995a, b). However, patients with other mycoses also test positive in these assays, and therefore the specificity of these tests is unsatisfactory (Miyazaki et al. 1995a, b). 17.12.2 Molecular Detection Sugita et al. (1999) developed an accurate identification system for all the species in the genus Trichosporon, including the six medically relevant species, based on comparative sequence analysis of the ITS regions. Furthermore, the same authors described a nested PCR assay for species-specific detection of T. asahii, which is the major causative agent for deep-seated infection (Sugita et al. 2001) (Table 17.3). Recently, DNA-based procedures for the identification of Trichosporon have been stepped up by the sequence analysis of the rRNA ITS1, which better distinguishes between sibling species (Table 17.3). Quantification of Trichosporon spp. in clinical samples is of great clinical significance. A quantitative real-time PCR assay to detect T. asahii DNA from the blood sample of thichosporonosis patients was recently developed (Mekha et al. 2007). The specific primer/probe system was capable of detecting T. asahii DNA in all samples from trichosporonosis patients, showing higher sensitivity than polysaccharide antigen detection (Table 17.3). 17.13 Zygomycetes Zygomycosis is an increasingly important IFI caused by emerging pathogens belonging to the Zygomycetes class, which can be subdivided into two orders: Mucorales and Entomophthorales (Ribes et al. 2000). The most common species of Mucorales causing angioinvasive zygomycosis are Rhizopus arrhizus (syn. Rhizopus oryzae), Rhizopus microsporus var. rhizopodiformis, and Rhizomucor pusillus. Other causative species include Absidia corymbifera, Mucor spp., and Cunninghamella bertholletiae (Chayakulkeeree et al. 2006). Human disease is most commonly caused by Mucorales, which are characterized by rapidly evolving course, tissue destruction, and invasion of blood vessels (Chayakulkeeree et al. 2006). Mycoses caused by Entomophthorales are more indolent and chronically progressive (Chayakulkeeree et al. 2006). The most common types of infection are sinus, pulmonary, and cutaneous; disseminated disease is reported in approximately one-fourth of patients (Roden et al. 2005). Iron overload predisposes to zygomycosis (Gonzalez et al. 2002). Mucorales are distributed worldwide, while Entomophthorales are generally limited to the tropics and subtropics (Gonzalez et al. 394 L. Putignani et al. 2002). Notably, there has been a steady increase in reported cases of zygomycosis over the last 60 years (Roden et al. 2005). Diabetic patients, particularly those with ketoacidosis, are highly susceptible to zygomycosis (Chayakulkeeree et al. 2006). Other risk factors are neutropenia, corticosteroid use, HSCT, SOT, HIV, drug use, skin or soft tissue breakdown (e.g., burn, traumatic inoculation, surgical wounds), prematurity, malnourishment, malignancy, and long-term prophylaxis with voriconazole. 17.13.1 Laboratory Diagnosis The traditional diagnosis of zygomycosis is based upon identification of broad, ribbon-like, pauciseptate hyphae by histopathology, or the use of macroscopic and microscopic morphology analysis following fungal culture (Gonzalez et al. 2002). Histopathology determinations suffer from subjectivity that is dependent upon the experience of the reader. In addition, tissue processing, fixation, and staining may require several days (Frater et al. 2001). When culture is used, the distinctive hyphal elements of zygomycetes are difficult to distinguish from those of other filamentous fungi, especially at early growth stages. Although the zygomycetes grow quite rapidly on solid media, sporulation and identification may still take several days. In addition, when a zygomycetous infection is not specifically suspected at the time of specimen submission, tissues may undergo extensive processing prior to culture, which ruptures the pauciseptate hyphae. It is therefore not unusual to have negative fungal culture results when histopathology results suggest the presence of zygomycosis (Kontoyiannis et al. 2000). 17.13.2 Molecular Detection DNA-based molecular techniques have an enormous potential for rapidly and accurately identifying the agents of zygomycoses. A molecular approach for the detection of Zygomycete molds may increase sensitivity and rapid diagnosis, resulting in earlier, directed therapy. Several reports describe utilization of universal fungal primers from the 18S rDNA or the ITS region for PCR amplification, followed by amplicon sequencing (Kobayashi et al. 2004; Schwarz et al. 2006) or hybridization to specific probes (Einsele et al. 1997, Rickerts et al. 2001). Also a seminested PCR assay targeting the 18S rDNA of Zygomycetes, followed by sequencing of the amplified product has been developed, providing sufficient sensitivity for fungal detection in paraffin-embedded tissues (Bialek et al. 2005a) (Table 17.3). A multiplex PCR method designed to yield different sized PCR products from the ITS region, whose length and combination correlates with four Rhizopus spp., provided detection of DNA from paraffin-embedded tissues and blood taken from a single patient (Nagao et al. 2005). Another assay uses a mixture of four genus-specific sense primers coupled with one degenerate antisense primer 17 DNA-Based Detection of Human Pathogenic Fungi 395 from the 18S rDNA to detect the four major genera within the Mucorales. RFLP of the amplicon makes possible species-level identification (Machouart et al. 2006). Kontoyiannis et al. (2005) used sequence-based identification to characterize the distribution of the various genera among clinical isolates of Zygomycetes. ITS sequencing identified the genus in almost all of their isolates (Kontoyiannis et al. 2005). They also showed that the results of ITS sequencing were > 20% discordant with those of morphological identification, and that most of the morphologically misidentified Zygomycetes isolates belonged to Rhizopus spp. (Kontoyiannis et al. 2005). More recent studies reported the use of real-time PCR assays for detection of infections caused by Zygomycetes. One study utilized a conserved region of the zygomycete cytochrome b gene as target (Hata et al. 2008). The second study was aimed at identifying the most common species of Zygomycetes causing human disease, using the 28S rRNA gene as target. The first assay (qPCR-1) detects Rhizopus, Mucor, and Rhizomucor spp. and the second assay (qPCR-2) detects Cunninghamella spp. Both assays utilize fluorescence resonance energy transfer (FRET) hybridization probes for detection of species within the genera (Kasai et al. 2008). 17.14 Dermatophytes The dermatophytes are a group of closely related fungi that have the ability to invade the stratum corneum of the epidermis and keratinized tissues (keratinophilic fungi), such as the skin, nail, and hair of humans and animals. Many dermatophytes are saprophytic soil fungi (geophilic), without any pathogenic feature for humans and animals and other only animals (zoophilic). Because zoophilic dermatophytes, through zoonotic transmission, infect humans only rarely and geophilic dermatophytes are infrequent cause of human disease, we will mainly focus on anthropophilic dermatophytes. These fungi cause dermathophytoses at most skin sites, although the feet, groin, scalp, and nails are most commonly affected. There are three genera of pathogenic anthropophilic dermatophytes, Trichophyton, Microsporum, and Epidermophyton. The last genus is represented by the single species Epidermophyton floccosum. Most of the 39 dermatophyte species are parasitic, causing disease in either humans and animals with mechanisms of narrow host range of adaptation (Hay 1995). The taxonomy of these fungi is complicated by the fact that most clinical isolates are imperfect fungi, organisms that do not produce sexual structures in culture. It is important for these fungi the strain differentiation within the same species to understand spreading of infections and relapse, quite frequent after apparently successful treatment (Abdel-Rahman 2008). Anthropophilic dermatophytes are the most common cause of human dermatophytosis (generally referred as ringworm or tinea corporis) as well as tinea pedis or tinea cruris, all of them belonging to the so-called superficial fungal infections (Kanbe 2008). Dermatomycoses are generally restricted to the cutaneous and the nonliving corneum layers due to host defense responses to the invading fungi in 396 L. Putignani et al. immunopotent individuals (Ogawa et al. 1998). In immunocompromised individuals, infection may progress to deep cutaneous and subcutaneous sites (Kanbe 2008). Dermatophytes are the most common causes of skin disease in tropical countries. Trichophyton rubrum is probably the most prevalent agent of dermatophytosis throughout the world (Veer et al. 2007). 17.14.1 Laboratory Diagnosis The morphological similarity, variability, and polymorphism of dermatophytes have meant that species identification for dermatophytes is time consuming and requires significant expertise. Furthermore, the application of chemotherapy has also contributed to the occasional modification and alteration of the morphological features of dermatophytes, resulting in atypical colonial growth and appearance and complicating laboratory identification procedures based on phenotypic features. 17.14.2 Molecular Detection Molecular analyzes of dermatophyte genomes would clear several problems in the traditional morphology-based taxonomy (Kanbe 2008). With these expectations, many investigators have focused their research on the nucleic acids of dermatophytes. The G + C content of chromosomal DNA isolated from species belonging to genera, Trichophyton, Microsporum, and Epidermophyton is 48.7–50.3%, a narrow range when compared with the 48–61% range reported in the single genus Aspergillus. This indicates a genetic homogeneity among dermatophytes as opposite to their phenotypic and ecological variation (Davidson et al. 1980). Subsequently many investigators have focused on mitochondrial DNA (mtDNA) and rDNA of dermatophytes to determine the phylogenetic relationships among dermatophytic species. These data have widely contributed to the development of molecular techniques for the identification and epidemiology of dermatophytes, as reviewed by Kanbe (2008), Blanz et al. (2000) and Kac (2000). PCR and PCRRFLP-based techniques made it possible to identify dermatophytes to species level, and to discriminate between isolates at the strain level (Makimura et al. 1999; Gräser et al. 2000; Liu et al. 2002). DNA samples obtained from fungal cultures or clinical specimens can be amplified by three primer sets targeting the 18S rDNA and ITS regions (TR1/TR2 and B2F/B4R for 18S rDNA and ITS1/ITS2 for ITS regions), yielding species-specific products from the ITS regions of Trichophyton, Microsporum, or Candida species (Turin et al. 2000). PCR-RFLP targeting both the ITS and NTS regions provides a valuable tool for species identification of dermatophytes (Jackson et al. 1999). The ITS is amplified with a universal primer set for 17 common dermatophyte species, and further RFLP with MvaI allows to discriminate between the majority of the dermatophytes at the species level, except for the sibling species of T. rubrum/Trichophyton soudanense and Trichophyton 17 DNA-Based Detection of Human Pathogenic Fungi 397 quinkeanum/Trichophyton schoenlenii. On the basis of RFLP profiles of the NTS region, T. rubrum can be subdivided into 14 types (Jackson et al. 1999). Similarly, Mochizuki et al. (2003) described the identification of T. tonsurans by PCR-RFLP targeting the ITS region, and MvaI or HinfI digestion. The RFLP profile obtained with MvaI was distinctive of T. tonsurans and different from that of other dermatophyte species (Mochizuki et al. 2003). Recently, Yoshida et al. (2006) reported on a PCR-based identification of T. tonsurans, in which the T. tonsurans genomic DNA was amplified by primers specific to the ITS1 region of this fungus. This method was able to discriminate T. tonsurans from other fungal species involved in dermatophytosis using only a single PCR amplification step (Yoshida et al. 2006). Real-time PCR with LightCycler has been combined with RFLP for the rapid detection/identification of isolates of dermatophytes and other pathogenic fungi in clinical specimens. Dermatophytes were distinguished from nondermatophytes using the LightCycler melting points. However, restriction enzyme digestion of the PCR product was necessary for discrimination between dermatophyte species (Gutzmer et al. 2006). Some other genes such as chitin synthase I gene (CHS1) or DNA topoisomerase II gene (TOP2) have also been used as a target for species identification of dermatophytes (Kano et al. 1999; Kanbe et al. 2003). 17.15 Conclusion and Perspectives Phenotype-based mycological diagnosis relying upon conventional microbiology and histopathology is time consuming, labor intensive, and often does not warrant sufficient sensitivity and specificity. Phenotype-based identification intrinsically suffers from the variable nature of phenotypic characteristics, which are influenced by culture conditions and subjective interpretation. These issues pose the need, among clinical mycologists, for faster, easier, and more proficient diagnostic tools for fungal infections. DNA-based methods have largely satisfied these needs because of their objective nature, rapidity, broad range of detection (panfungal tests), applicability to a variety of specimen types, capability to provide results that are unaffected by growth conditions, and high discriminatory power, thereby ensuring identification of fungi that would be indistinguishable according to phenotypic traits. Our review of the current literature on DNA-based methods for fungal identification has highlighted the utility of such methods in clinical practice. In many instances, the simple PCR and sequencing of conserved regions of the fungal genome provides species-level identification with reasonable confidence for both molds and yeasts, including those species which escape identification by conventional techniques or commercially available kits. Furthermore, molecular assays are becoming the only suitable diagnostic approach to early detection of fungal cells (i.e., of their DNA) at low infecting concentrations and in the absence of serological evidence, as in case of immunocompromised patients (White et al. 2006; Badiee et al. 2009). Early diagnosis will allow clinicians to combat fungal infection at an 398 L. Putignani et al. early stage through choice-specific and effective treatment, avoiding empirical therapy and development of resistance to antimycotic drugs. In this scenario, it is expected that ongoing progress in molecular biology will continue to have a positive impact on medical mycology. There are many issues, however, that need to be addressed in the future. First, the direct identification of fungal DNA extracted from pathological samples still poses some problem, and most molecular identification methods require preliminary isolation fungi as pure cultures. Second, the percentage of sequence homology required to define fungal species or genera should be clearly established by the scientific community for the main phylogenetic markers, e.g., the rRNA genes. Third, interrogation of public databases of rRNA genes can provide dubious identities because of the presence of uncontrolled entries. If fungal cultures are unavailable, uncertain, or spurious, DNA-based identifications cannot be verified through appropriate phenotypic testing, and the molecular epidemiology of infectious fungi will be difficult to trace. Fourth, the reliability of the current DNA-based identification assays ought to be assessed in the coming years with well-designed prospective studies and regular quality controls. Rigorous testing of molecular assays for the diagnosis of fungal disease through multicenter proficiency studies and external validation from certifying institutions (e.g., USFDA) will be unavoidable steps. Lastly, the clinical meaning of a positive PCR should be test interpreted in the light of our knowledge of the clearance kinetics of fungal DNA in the infected patient; this would help in distinguishing between colonized individuals and infected patients. Simplification and/or standardization of methods for DNA extraction and detection are likely to facilitate the introduction of molecular technology to routine clinical laboratories (De Marco et al. 2007). But technical problems still need to be solved, primarily the risk of contamination. Nonetheless, the approach of using primers that target multicopy genes is likely to provide the high degrees of sensitivity that are needed for initial diagnosis of severe fungal infection. Automated DNA extraction, real-time PCR techniques, and the development of controlled sequence databases are likely to improve reliability of the current DNAbased identification assays. The usage of close real-time PCR systems (e.g., Light Cycler SeptiFast) in the current diagnostic protocols and the development of repPCR-based methods (e.g., DiversiLab) for further routine evaluation are partly overcoming contamination issues. At the same time, these commercial systems provide ready-to-use instruments and reagents that render molecular diagnosis affordable even in non specialized laboratories. New advanced molecular approaches based on mass spectrometry (e.g., MALDI-ToF, Matrix-assisted laser description ionization-time of flight-mass spectrometry) are currently under setting for the rapid detection and identification of fungi in clinical microbiology laboratories. They could open new horizons for the fast and reliable management of fungal infections in high risk patients. Acknowledgments This work was supported by Grants from: Ricerca Corrente, Bambino Gesù Hospital Health Care and Research Institute RC 200702P002153 to L.P.; Ricerca Corrente 2008, 17 DNA-Based Detection of Human Pathogenic Fungi 399 INMI “Lazzaro Spallanzani” to P.V.; ISPESL, Ricerca Finalizzata PMS/40/06 to P.V. We acknowledge with great respect the work of very many researchers that has contributed to improve molecular diagnosis of mycoses, and that for reasons of space we were not able to directly reference here. References Abdel-Rahman SM (2008) Strain differentiation of dermatophytes. Mycopathologia 166:319–333 Badiee P, Kordbacheh P, Albori A, Malekhoseini SA (2007) Invasive fungal infection in renal transplant recipients demonstrated by panfungal polymerase chain reaction. Exp Clin Transplant 5:624–629 Badiee P, Kordbacheh P, Alborzi A, Malekhoseini S, Ramzi M, Mirhendi H, Mahmoodi M, Shakiba E (2009) Study on invasive fungal infections in immunocompromised patients to present a suitable early diagnostic procedure. Int J Infect Dis 13:97–102 Bartie KL, Williams DW, Wilson MJ, Potts AJ, Lewis MA (2001) PCR fingerprinting of Candida albicans associated with chronic hyperplastic candidosis and other oral conditions. J Clin Microbiol 39:4066–4075 Basková L, Landlinger C, Preuner S, Lion T (2007) The Pan-AC assay: a single-reaction real-time PCR test for quantitative detection of a broad range of Aspergillus and Candida species. J Med Microbiol 56:1167–1173 Beretta S, Fulgencio JP, Enache-Angoulvant A, Bernard C, El Metaoua S, Ancelle T, Denis M, Hennequin C (2006) Application of microsatellite typing for the investigation of a cluster of cases of Candida albicans candidaemia. Clin Microbiol Infect 12:674–676 Bergman A, Fernandez V, Holmström KO, Claesson BE, Enroth H (2007) Rapid identification of pathogenic yeast isolates by real-time PCR and two-dimensional melting-point analysis. Eur J Clin Microbiol Infect Dis 26:813–818 Bialek R, Fischer J, Feucht A, Najvar LK, Dietz K, Knobloch J, Graybill JR (2001) Diagnosis and monitoring of murine histoplasmosis by a nested PCR assay. J Clin Microbiol 39:1506–1509 Bialek R, Feucht A, Aepinus C, Just-Nübling G, Robertson VJ, Knobloch J, Hohle R (2002a) Evaluation of two nested PCR assays for detection of Histoplasma capsulatum DNA in human tissue. J Clin Microbiol 40:1644–1647 Bialek R, Weiss M, Bekure-Nemariam K, Najvar LK, Alberdi MB, Graybill JR, Reischl U (2002b) Detection of Cryptococcus neoformans DNA in tissue samples by nested and real-time PCR assays. Clin Diagn Lab Immunol 9:461–469 Bialek R, Cirera AC, Herrmann T, Aepinus C, Shearn-Bochsler VI, Legendre AM (2003) Nested PCR assays for detection of Blastomyces dermatitidis DNA in paraffin-embedded canine tissue. J Clin Microbiol 41:205–208 Bialek R, Kern J, Herrmann T, Tijerina R, Cecenas L, Reischl U, Gonzalez GM (2004) PCR assays for identification of Coccidioides posadasii based on the nucleotide sequence of the antigen2/ proline-rich antigen. J Clin Microbiol 42:778–783 Bialek R, González GM, Begerow D, Zelck UE (2005a) Coccidioidomycosis and blastomycosis: advances in molecular diagnosis. FEMS Immunol Med Microbiol 45:355–360 Bialek R, Konrad F, Kern J, Aepinus C, Cecenas L, Gonzalez GM, Just-Nübling G, Willinger B, Presterl E, Lass-Flörl C, Rickerts V (2005b) PCR based identification and discrimination of agents of mucormycosis and aspergillosis in paraffin wax embedded tissue. J Clin Pathol 58:1180–1184 Bille J, Marchetti O, Calandra T (2005) Changing face of health-care associated fungal infections. Curr Opin Infect Dis 18:314–319 400 L. Putignani et al. Binnicker MJ, Buckwalter SP, Eisberner JJ, Stewart RA, McCullough AE, Wohlfiel SL, Wengenack NL (2007) Detection of Coccidioides species in clinical specimens by real-time PCR. J Clin Microbiol 45:173–178 Birch PRJ, Sims PFG, Broda P (1992) Nucleotide sequence of a gene from Phanerochaete chrysosporium that shows homology to the facA gene of Aspergillus nidulans. DNA Sequence 2:319–323 Bishop JA, Chase N, Lee R, Kurtzman CP, Merz WG (2008) Production of white colonies on CHROMagar Candida medium by members of the Candida glabrata clade and other species with overlapping phenotypic traits. J Clin Microbiol 46:3498–3500 Blair JE, Logan JL (2001) Coccidioidomycosis in solid organ transplantation. Clin Infect Dis 33:1536–1544 Blanz P, Buzina W, Ginter G, Gräser Y (2000) Molecular biological methods in taxonomy and diagnostics of dermatophytes. Mycoses 43:11–16 Blumberg HM, Jarvis WR, Soucie JM, Edwards JE, Patterson JE, Pfaller MA, Rangel-Frausto MS, Rinaldi MG, Saiman L, Wiblin RT, Wenzel RP, The NEMIS Study Group (2001) Risk factors for candidal bloodstream infections in surgical intensive care unit patients: the NEMIS prospective multicenter study. Clin Infect Dis 33:177–186 Bougnoux ME, Dupont C, Mateo J, Saulnier P, Faivre V, Payen D, Nicolas-Chanoine MH (1999) Serum is more suitable than whole blood for diagnosis of systemic candidiasis by nested PCR. J Clin Microbiol 37:925–930 Boutati EI, Anaissie EJ (1997) Fusarium, a significant emerging pathogen in patients with hematologic malignancy: ten years’ experience at a cancer center and implications for management. Blood 90:999–1008 Bowen AR, Chen-Wu JL, Momany M, Young R, Szaniszlo PJ, Robbins PW (1992) Classification on fungal chitin synthases. Proc Natl Acad Sci USA 89:519–523 Boyanton BL Jr, Luna RA, Fasciano LR, Menne KG, Versalovic J (2008) DNA pyrosequencingbased identification of pathogenic Candida species by using the internal transcribed spacer 2 region. Arch Pathol Lab Med 132:667–674 Bracca A, Tosello ME, Girardini JE, Amigot SL, Gomez C, Serra E (2003) Molecular detection of Histoplasma capsulatum var capsulatum in human clinical samples. J Clin Microbiol 41:1753–1755 Bretagne S (2003) Molecular diagnostics in clinical parasitology and mycology: limits of the current polymerase chain reaction (PCR) assays and interest of the real-time PCR assays. Clin Microbiol Infect 9:505–511 Bretagne S, Costa JM (2005) Towards a molecular diagnosis of invasive aspergillosis and disseminated candidosis. FEMS Immunol Med Microbiol 45:361–368 Bretagne S, Costa JM, Marmorat-Khuong A, Poron F, Cordonnier C, Vidaud M, Fleury-Feith J (1995) Detection of Aspergillus species DNA in bronchoalveolar lavage samples by competitive PCR. J Clin Microbiol 33:1164–1168 Bretagne S, Costa JM, Bart-Delabesse E, Dhédin N, Rieux C, Cordonnier C (1998) Comparison of serum galactomannan antigen detection and competitive polymerase chain reaction for diagnosing invasive aspergillosis. Clin Infect Dis 26:1407–1412 Bruns TD, Vilgalys R, Barns SM, Gonzalez D, Hibbett DS, Lane DJ, Simon L, Stickel S, Szaro TM, Weisburg WG, Sogin ML (1992) Evolutionary relationships within the fungi: analyses of small subunit ribosomal DNA sequences. Appl Environ Microbiol 61:681–689 Buchheidt D, Hummel M (2005) Aspergillus polymerase chain reaction (PCR) diagnosis. Med Mycol 43:S139–S145 Buchman TG, Rossier M, Mertz WG, Charache P (1990) Detection of surgical pathogens by in vitro DNA amplification Part I Rapid identification of Candida albicans by in vitro amplification of a fungus specific gene. Surgery 108:338–347 Bullok WE (1995) Histoplasma capsulatum. In: Mandell GL, Bennet J, Dolin R (eds) Principles and practice of infectious diseases, vol 1. Elsevier Churchill Livingstone, New York, NY, pp 2340–2353 17 DNA-Based Detection of Human Pathogenic Fungi 401 Burgener-Kairuz P, Zuber JP, Jaunin P, Buchman TG, Bille J, Rossier MJ (1994) Rapid detection and identification of Candida albicans and Torulopsis (Candida) glabrata in clinical specimens by species-specific nested PCR amplification of a cytochrome P-450 lanosterol-alphademethylase (L1A1) gene fragment. Clin Microbiol 32:1902–1907 Burnie JP, Golband N, Matthews RC (1997) Semiquantitative polymerase chain reaction enzyme immunoassay for diagnosis of disseminated candidiasis. Eur J Clin Microbiol Infect Dis 16:346–350 Cairns L, Blythe D, Kao A, Pappagianis D, Kaufman L, Kobayashi J, Hajjeh R (2000) Outbreak of coccidioidomycosis in Washington State residents returning from Mexico. Clin Infect Dis 30:61–64 Chagas-Neto TC, Chaves GM, Colombo AL (2008) Update on the genus Trichosporon. Mycopathologia 166:121–132 Challier S, Boyer S, Abachin E, Berche P (2004) Development of a serum-based Taqman real-time PCR assay for diagnosis of invasive aspergillosis. J Clin Microbiol 42:844–846 Chang HC, Leaw SN, Huang AH, Wu TL, Chang TC (2001) Rapid identification of yeasts in positive blood cultures by a multiplex PCR method. J Clin Microbiol 39:3466–3471 Chang DC, Anderson S, Wannemuehler K, Engelthaler DM, Erhart L, Sunenshine RH, Burwell LA, Park BJ (2008) Testing for coccidioidomycosis among patients with community-acquired pneumonia. Emerg Infect Dis 14:1053–1059 Chayakulkeeree M, Ghannoum MA, Perfect JR (2006) Zygomycosis: the re-emerging fungal infection. Eur J Clin Microbiol Infect Dis 25:215–229 Chen SC, Halliday CL, Meyer W (2002) A review of nucleicacid-based diagnosis tests for systemic mycoses with an emphasis on polymerase chain reaction-based assays. Med Mycol 40:333–357 Chen SC, Tong ZS, Lee OC, Halliday C, Playford EG, Widmer F, Kong FR, Wu C, Sorrell TC (2008) Clinician response to Candida organisms in the urine of patients attending hospital. Eur J Clin Microbiol Infect Dis 27:201–208 Chiller TM, Galgiani JN, Stevens DA (2003) Coccidioidomycosis. Infect Dis Clin North Am 17:41–57 Chowdhary A, Guarro J, Randhawa HS, Gené J, Cano J, Jain RK, Kumar S, Khanna G (2008) A rare case of chromoblastomycosis in a renal transplant recipient caused by a non-sporulating species of Rhytidhysteron. Med Mycol 46:163–166 Chryssanthou E, Andersson B, Petrini B, Lofdahl S, Tollemar J (1994) Detection of Candida albicans DNA in serum by polymerase chain reaction. Scand J Infect Dis 26:479–485 Cornely OA (2008) Aspergillus to Zygomycetes: causes, risk factors, prevention, and treatment of invasive fungal infections. Infection 36:296–313 Costa C, Vidaud D, Olivi M, Bart-Delabesse E, Vidaud M, Bretagne S (2001) Development of two real-time quantitative TaqMan PCR assays to detect circulating Aspergillus fumigatus DNA in serum. J Microbiol Methods 44:263–269 Costa C, Costa JM, Desterke C, Botterel F, Cordonnier C, Bretagne S (2002) Real-time PCR coupled with automated DNA extraction and detection of galactomannan antigen in serum by enzyme-linked immunosorbent assay for diagnosis of invasive aspergillosis. J Clin Microbiol 40:2224–2227 Cox GM, Rude TH, Dykstra CC, Perfect JR (1995) The actin gene from Cryptococcus neoformans: structure and phylogenetic analysis. J Med Vet Mycol 33:261–266 Crampin AC, Matthews RC (1993) Application of the polymerase chain reaction to the diagnosis of candidosis by amplification of an HSP 90 gene fragment. J Med Microbiol 39:233–238 Cuenca-Estrella M, Meije Y, Diaz-Pedroche C, Gomez-Lopez A, Buitrago MJ, Bernal-Martinez L, Grande C, Juan RS, Lizasoain M, Rodriguez-Tudela JL, Aguado JM (2009) Value of serial quantification of fungal DNA by a real-time PCR-based technique for early diagnosis of invasive Aspergillosis in patients with febrile neutropenia. J Clin Microbiol 47:379–384 D’Antonio D, Violante B, Mazzoni A, Bonfini T, Capuani MA, D’Aloia F, Lacone A, Schioppa F, Romano F (1998) A nosocomial cluster of Candida inconspicua infections in patients with haematological malignancies. J Clin Microbiol 36:792–795 402 L. Putignani et al. Dalle F, Lopez J, Caillot D, Cuisenier B, Ecarnot LA, Dumont L, Bonnin A (2002) False-positive results caused by cotton swabs in commercial Aspergillus antigen latex agglutination test. Eur J Clin Microbiol Infect Dis 21:130–132 Davidson FD, Mackenzie DWR, Owen RJ (1980) Deoxyribonucleic acid base compositions of dermatophytes. J Gen Microbiol 118:465–470 de Aguirre L, Hurst SF, Choi JS, Shin JH, Hinrikson HP, Morrison CJ (2004) Rapid differentiation of Aspergillus species from other medically important opportunistic molds and yeasts by PCRenzyme immunoassay. J Clin Microbiol 42:3495–3504 de Hoog GS, Gerrits van den Ende AHG (1992) Nutritional pattern and eco-physiology of Hortaea werneckii, agent of human tinea nigra. Antonie Van Leeuwenhoek Int J G 62:321–329 de Hoog GS, Gerrits van den Ende AHG (1998) Molecular diagnostics of clinical strains of filamentous Basidiomycetes. Mycoses 41:183–189 De Marco D, Perotti M, Ossi CM, Burioni R, Clementi M, Mancini N (2007) Development and validation of a molecular method for the diagnosis of medically important fungal infections. New Microbiol 30:308–312 Diekema DJ, Pfaller MA (2004) Nosocomial candidemia: an ounce of prevention is better than a pound of cure. Infect Control Hosp Epidemiol 25:624–626 Dunyach C, Bertout S, Phelipeau C, Drakulovski P, Reynes J, Mallié M (2008) Diagn Detection and identification of Candida spp in human serum by Light Cycler real-time polymerase chain reaction. Microbiol Infect Dis 60:263–271 Edwards JE (1991) Invasive Candida infections. N Engl J Med 324:1060–1062 Eggimann P, Francioli P, Bille J, Schneider R, Wu M-W, Chapuis G, Chiolero R, Pannatier A, Schilling J, Geroulanos S, Glauser MP, Calandra T (1999) Fluconazole prophylaxis prevents intraabdominal candidiasis in high-risk surgical patients. Crit Care Med 27:1066–1072 Einsele H, Hebart H, Roller G, Loffler J, Rothenhofer I, Muller CA, Bowden RA, van Burik J, Engelhard D, Kanz L, Schumacher U (1997) Detection and identification of fungal pathogens in blood by using molecular probes. J Clin Microbiol 35:1353–1360 Elie CM, Lott TJ, Reiss E, Morrison CJ (1998) Rapid identification of Candida species with species-specific probes. J Clin Microbiol 36:3260–3265 Ersoz G, Otag F, Erturan Z, Aslan G, Kaya A, Emekdas G, Sugita T (2004) An outbreak of Dipodascus capitatus infection in the ICU: three case reports and review of the literature. Jpn J Infect Dis 57:248–252 Evertsson U, Monstein HJ, Johansson AG (2000) Detection and identification of fungi in blood using broad-range 28S rRNA PCR amplification and species-specific hybridisation. APMIS 108:385–392 Faber J, Moritz N, Henninger N, Zepp F, Knuf M (2009) Rapid detection of common pathogenic Aspergillus species by a novel real-time PCR approach. Mycoses 52:228–233 Farr RC, Gardner G, Acker JD, Brint JM, Haglund LF, Land M, Schweitzer JB, West BC (1992) Blastomycotic cranial osteomyelitis. Am J Otol 13:582–586 Fisher MC, Koenig GL, White TJ, San-Blas G, Negroni R, Alvarez IG, Wanke B, Taylor JW (2001) Biogeographic range expansion into South America by Coccidioides immitis mirrors New World patterns of human migration. Proc Natl Acad Sci U S A 98:4558–62 Fisher MC, Koenig GL, White TJ, Taylor JW (2002) Molecular and phenotypic description of Coccidioides posadasii sp nov, previously recognized as the non-California population of Coccidioides immitis. Mycologia 94:73–84 Franzot SP, Salkin IF, Casadevall A (1999) Cryptococcus neoformans var grubii: separate varietal status for Cryptococcus neoformans serotype A isolates. J Clin Microbiol 37:838–840 Frater JL, Hall GS, Procop GW (2001) Histologic features of zygomycosis: emphasis on perineural invasion and fungal morphology. Arch Pathol Lab Med 125:375–378 Fredricks DN, Smith C, Meier AJ (2005) Comparison of six DNA extraction methods for recovery of fungal DNA as assessed by quantitative PCR. Clin Microbiol 43:5122–5128 Freifeld AG, Iwen PC, Lesiak BL, Gilroy RK, Stevens RB, Kalil AC (2005) Histoplasmosis in solid organ transplant recipients at a large Midwestern university transplant center. Transpl Infect Dis 7:109–115 17 DNA-Based Detection of Human Pathogenic Fungi 403 Friedman GD, Jeffrey Fessel W, Udaltsova NV, Hurley LB (2005) Cryptococcosis: the 1981–2000 epidemic. Mycoses 48:122–125 Galgiani JN (1992) Coccidioidomycosis: changes in clinical expression, serological diagnosis, and therapeutic options. Clin Infect Dis 14:S100–S105 Galgiani JN (1999) Coccidioidomycosis: a regional disease of national importance – rethinking approaches for control. Ann Intern Med 130:293–300 Geha DJ, Roberts GD (1994) Laboratory detection of fungemia. Clin Lab Med 14:83–97 Girmenia C, Martino P, De Bernardis F, Cassone A (1997) Assessment of detection of Candida mannoproteinemia as a method to differentiate central venous catheter-related candidemia from invasive disease. J Clin Microbiol 35:903–906 Girmenia C, Pagano L, Martino B, D’Antonio D, Fanci R, Specchia G, Melillo L, Buelli M, Pizzarelli G, Venditti M, Martino P, Infection Program GIMEMA (2005) Invasive infections caused by Trichosporon species and Geotrichum capitatum in patients with hematological malignancies: a retrospective multicenter study from Italy and review of the literature. J Clin Microbiol 43:1818–1828 Gomez BL, Figueroa JI, Hamilton AJ, Ortiz BL, Robledo MA, Restrepo A, Hay RJ (1997) Development of a novel antigen detection test for histoplasmosis. J Clin Microbiol 35:2618–2622 Gonzalez CE, Rinaldi MG, Sugar AM (2002) Zygomycosis. Infect Dis Clin North Am 16:895–914 Gräser Y, Kuijpers AFA, Presber W, de Hoog GS (2000) Molecular taxonomy of the Trichophyton rubrum complex. J Clin Microbiol 38:3329–3336 Griffin DW, Kellogg CA, Peak KK, Shinn EA (2002) A rapid and efficient assay for extracting DNA from fungi. Lett Appl Microbiol 34:210–214 Griffiths LJ, Anyim M, Doffman SR, Wilks M, Millar MR, Agrawal SG (2006) Comparison of DNA extraction methods for Aspergillus fumigatus using real-time PCR. J Med Microbiol 55:1187–1191 Grigis A, Farina C, Symoens F, Nolard N, Goglio A (2000) Nosocomial pseudo-outbreak of Fusarium verticillioides associated with sterile plastic containers. Infect Control Hosp Epidemiol 21:50–52 Gromadzki SG, Chaturvedi V (2000) Limitation of the AccuProbe Coccidioides immitis culture identification test: falsenegative results with formaldehyde-killed cultures. J Clin Microbiol 38:2427–2428 Guarro J, Gené J, Stchigel AM (1999) Developments in fungal taxonomy. Clin Microbiol Rev 12:454–500 Guedes HL, Guimarães AJ, Muniz Mde M, Pizzini CV, Hamilton AJ, Peralta JM, Deepe GS Jr, Zancopé-Oliveira RM (2003) PCR assay for identification of Histoplasma capsulatum based on the nucleotide sequence of the M antigen. J Clin Microbiol 41:535–539 Guého E, Improvisi L, Christen R, de Hoog GS (1993) Phylogenetic relationships of Cryptococcus neoformans and some related basidiomycetous yeasts determined from partial large subunit rRNA sequences. Antonie Van Leeuwenhoek Int J Genet 63:175–189 Guiver M, Levi K, Oppenheim BA (2001) Rapid identification of Candida species by TaqMan PCR. J Clin Pathol 54:362–366 Gupta AK, Boekhout T, Theelen B, Summerbell R, Batra R (2004) Identification and typing of Malassezia species by amplified fragment length polymorphism and sequence analyses of the internal transcribed spacer and large-subumit regions of ribosomal DNA. J Clin Microbiol 42:4253–4260 Gutzmer R, Mommert S, Küttler U, Werfel T, Kapp A (2006) Rapid identification and differentiation of fungal DNA in dermatological specimens by LightCycler PCR. J Med Microbiol 53:1207–1214 Hall L, Wohlfiel S, Roberts GD (2003) Experience with the MicroSeq D2 large-subunit ribososmal DNA sequencing kit for identification of commonly encountered, clinically important yeast species. J Clin Microbiol 41:5099–5102 Hall L, Wohlfiel S, Roberts GD (2004) Experience with the MicroSeq D2 large-subunit ribosomal DNA sequencing kit for identification of filamentous fungi encountered in the clinical laboratory. J Clin Microbiol 42:622–626 404 L. Putignani et al. Halliday C, Wu QX, James G, Sorrell T (2005) Development of a nested qualitative real-time PCR assay to detect Aspergillus species DNA in clinical specimens. J Clin Microbiol 43:5366–5368 Hamilton JR, Noble A, Denning DW, Stevens DA (1991) Performance of cryptococcus antigen latex agglutination kits on serum and cerebrospinal fluid specimens of AIDS patients before and after pronase treatment. J Clin Microbiol 29:333–339 Hansen D, Healy M, Reece K, Smith C, Woods GL (2008) Repetitive-sequence-based PCR using the DiversiLab system for identification of Aspergillus species. J Clin Microbiol 46:1835–1839 Harmsen MC, Schuren FHJ, Moukha SM, van Zuilen CM, Punt PJ, Wessels JGH (1992) Sequence analysis of the glyceraldehyde-3-phosphate dehydrogenase genes from the basidiomycetes Schizophyllum commune, Phanerochaete chrysosporium and Agaricus bisporus. Curr Genet 22:447–454 Hashiguchi K, Niki Y, Soejima R (1994) Cyclophosphamide induces false positive results in detection of Aspergillus antigen in urine. Chest 105:975–976 Hassouna N, Michot B, Bachelleire J (1984) The complete nucleotide sequence of mouse 28S rRNA gene Implications for the process of size increase of the large subunit rRNA in higher eukaryotes. Nucleic Acids Res 8:3563–3583 Hata DJ, Buckwalter SP, Pritt BS, Roberts GD, Wengenack NL (2008) Real-time PCR method for detection of zygomycetes. J Clin Microbiol 46:2353–2358 Hay RJ (1995) Dermathophytosis and other superficial mycoses. In: Mandell GL, Bennet J, Dolin R (eds) Principles and practice of infectious diseases, vol 1. Elsevier Churchill Livingstone, New York, NY, pp 3051–3062 Hayden RT, Qian X, Roberts GD, Lloyd RV (2001) In situ hybridization for the identification of yeastlike organisms in tissue section. Diagn Mol Pathol 10:15–23 Hazen KC (1995) New and emerging yeast pathogens. Clin Microbiol Rev 8:462–478 Hazen KC (1996) Methods for fungal identification in the clinical mycology laboratory. Clin Microbiol News 18:137–144 Healy M, Reece K, Walton D, Huong J, Shah K, Kontoyiannis DP (2004) Identification to the species level and differentiation between strains of Aspergillus clinical isolates by automated repetitive-sequence-based PCR. J Clin Microbiol 42:4016–4024 Hendolin PH, Paulin L, Koukila-Kähkölä P, Anttila VJ, Malmberg H, Richardson M, Ylikoski J (2000) Panfungal PCR and multiplex liquid hybridization for detection of fungi in tissue specimens. J Clin Microbiol 38:4186–4192 Henry T, Iwen PC, Hinrichs SH (2000) Identification of Aspergillus species using internal transcribed spacer regions 1 and 2. J Clin Microbiol 38:1510–1515 Hibbett DS (1992) Ribosomal RNA and fungal systematics. Trans Mycol Soc Jpn 33:533–556 Hibbett DS, Binder M, Bischoff JF, Blackwell M, Cannon PF, Eriksson OE, Huhndorf S, James T, Kirk PM, Lücking R, Thorsten Lumbsch H, Lutzoni F, Matheny PB, McLaughlin DJ, Powell MJ, Redhead S, Schoch CL, Spatafora JW, Stalpers JA, Vilgalys R, Aime MC, Aptroot A, Bauer R, Begerow D, Benny GL, Castlebury LA, Crous PW, Dai YC, Gams W, Geiser DM, Griffith GW, Gueidan C, Hawksworth DL, Hestmark G, Hosaka K, Humber RA, Hyde KD, Ironside JE, Kõljalg U, Kurtzman CP, Larsson KH, Lichtwardt R, Longcore J, Miadlikowska J, Miller A, Moncalvo JM, Mozley-Standridge S, Oberwinkler F, Parmasto E, Reeb V, Rogers JD, Roux C, Ryvarden L, Sampaio JP, Schüssler A, Sugiyama J, Thorn RG, Tibell L, Untereiner WA, Walker C, Wang Z, Weir A, Weiss M, White MM, Winka K, Yao YJ, Zhang N (2007) A higher-level phylogenetic classification of the fungi. Mycol Res 111:509–547 Hinrikson HP, Hurst SF, Lott TJ, Warnock DW, Morrison CJ (2005) Assessment of ribosomal large-subunit D1–D2, internal transcribed spacer 1, and internal transcribed spacer 2 regions as targets for molecular identification of medically important Aspergillus species. J Clin Microbiol 43:2092–2103 Hockey LJ, Fujita NK, Gibson TR, Rotrosen D, Montgomerie JZ, Edwards JE Jr (1982) Detection of fungemia obscured by concomitant bacteremia: in vitro and in vivo studies. J Clin Microbiol 16:1080–1085 17 DNA-Based Detection of Human Pathogenic Fungi 405 Hope WW, Walsh TJ, Denning DW (2005) Laboratory diagnosis of invasive aspergillosis. Lancet Infect Dis 5:609–622 Hopfer RL, Walden P, Setterquist S, Highsmith WE (1993) Detection and differentiation of fungi in clinical specimens using polymerase chain reaction (PCR) amplification and restriction enzyme analysis. J Med Vet Mycol 31:65–75 Hospenthal DR (1995) Agents of chromoblastomycosis. In: Mandell GL, Bennet J, Dolin R (eds) Principles and practice of infectious diseases, vol 1. Elsevier Churchill Livingstone, New York, NY, pp 2988–2991 Hull CM, Heitman J (2002) Genetics of Cryptococcus neoformans. Annu Rev Genet 36:557–615 Inácio J, Flores O, Spencer-Martins I (2008) Efficient identification of clinically relevant Candida yeast species by use of an assay combining panfungal loop-mediated isothermal DNA amplification with hybridization to species-specific oligonucleotide probes. J Clin Microbiol 46:713–720 Innings A, Ullberg M, Johansson A, Rubin CJ, Noreus N, Isaksson M, Herrmann B (2007) Multiplex real-time PCR targeting the RNase P RNA gene for detection and identification of Candida species in blood. J Clin Microbiol 45:874–880 Ito-Kuwa S, Nakamura K, Aoki S, Vidotto V (2007) Serotype identification of Cryptococcus neoformans by multiplex PCR. Mycoses 50:277–281 Iwen PC (2003) Molecular detection and typing of fungal pathogens. Clin Lab Med 23:781–799 Iwen PC, Freifeld AG, Bruening TA, Hinrichs SH (2004) Use of a panfungal PCR assay for detection of fungal pathogens in a commercial blood culture system. J Clin Microbiol 42:2292–2293 Iyer RS, Banker DD (2002) Cryptococcal meningitis in AIDS. Indian J Med Sci 56:593–597 Jackson CJ, Barton RC, Evance EGV (1999) Species identification and strain differentiation of dermatophyte fungi by analysis of ribosomal-DNA intergenic spacer regions. J Clin Microbiol 37:931–936 Jordan JA (1994) PCR identification of four medically important Candida species by using a single primer pair. J Clin Microbiol 32:2962–2967 Kac G (2000) Molecular approaches to the study of dermatophytes. Med Mycol 38:329–336 Kami M, Fukui T, Ogawa S, Kazuyama Y, Machida U, Tanaka Y, Kanda Y, Kashima T, Yamazaki Y, Hamaki T, Mori S, Akiyama H, Mutou Y, Sakamaki H, Osumi K, Rimura S, Hirai H (2001) Use of real-time PCR on blood samples for diagnosis of invasive aspergillosis. Clin Infect Dis 33:1504–1512 Kan VL (1993) Polymerase chain reaction for the diagnosis of candidemia. J Infect Dis 168:779–783 Kanbe T (2008) Molecular approaches in the diagnosis of dermatophytosis. Mycopathologia 166:307–317 Kanbe T, Suzuki Y, Kamiya A, Mochizuki T, Fujihiro M, Kikuchi A (2003) PCR-based identification of common dermatophyte species using primer sets specific for the DNA topoisomerase II genes. J Dermatol Sci 32:151–161 Kano R, Nakamura Y, Watari T, Watanabe S, Takahashi H, Tsujimoto H, Hasegawa A (1999) Species-specific primers of chitin synthase 1 gene for the differentiation of the Trichophyton mentagrophytes complex. Mycoses 42:71–74 Kano R, Fujino Y, Takamoto N, Tsujimoto H, Hasegawa A (2001) PCR detection of the Cryptococcus neoformans CAPS9 gene from a biopsy specimen from a case of feline cryptococcosis. J Vet Diagn Invest 13:439–442 Kantarcioglu AS, de Hoog GS (2004) Infections of the central nervous system by melanized fungi: a review of cases presented between1999 and 2004. Mycoses 47:4–13 Kappe R, Rimek D (1999) Laboratory diagnosis of Aspergillus fumigatus-associated diseases. Contrib Microbiol 2:88–104 Kappe R, Schulze-Berge A (1993) New cause for false-positive results with the Pastorex Aspergillus antigen latex agglutination test. J Clin Microbiol 31:2489–2490 Karabinis A, Hill C, Leclerq B, Tancrède C, Baume D, Andremont A (1988) Risk factors for candidemia in cancer patients: a case-control study. J Clin Microbiol 26:429–432 406 L. Putignani et al. Kasai M, Harrington SM, Francesconi A, Petraitis V, Petraitiene R, Beveridge MG, Knudsen T, Milanovich J, Cotton MP, Hughes J, Schaufele RL, Sein T, Bacher J, Murray PR, Kontoyiannis DP, Walsh TJ (2008) Detection of a Molecular Biomarker for Zygomycetes by Quantitative PCR Assays in Plasma, Bronchoalveolar Lavage, and Lung Tissue in Pulmonary Zygomycosis. J Clin Microbiol 46:3690–3702 Kauffman CA (2006) Endemic mycoses: blastomycosis, histoplasmosis, and sporotrichosis. Infect Dis Clin North Am 20:645–662 Kaufman L, Valero G, Padhye AA (1998) Misleading manifestations of Coccidioides immitis in vivo. J Clin Microbiol 36:3721–3723 Kendrick B (1992) The fifth kingdom, 2nd edn. Mycologue Publications, Waterloo, Canada Kiska DL, Orkiszewski DR, Howell D, Gilligan PH (1994) Evaluation of new monoclonal antibody-based latex agglutination test for detection of cryptococcal polysaccharide antigen in serum and cerebrospinal fluid. J Clin Microbiol 32:2309–2311 Klingspor L, Jalal S (2006) Molecular detection and identification of Candida and Aspergillus spp from clinical samples using real-time PCR. Clin Microbiol Infect 12:745–753 Klingspor L, Loeffler J (2009) Aspergillus PCR formidable challenges and progress. Med Mycol 47:S241–247 Kobayashi M, Togitani K, Machida H, Uemura Y, Ohtsuki Y, Taguchi H (2004) Molecular polymerase chain reaction diagnosis of pulmonary mucormycosis caused by Cunninghamella bertholletiae. Respirology 9:397–401 Kontoyiannis DP, Bodey GP (2002) Invasive aspergillosis in 2002: an update. Eur J Clin Microbiol Infect Dis 21:161–172 Kontoyiannis DP, Wessel VC, Bodey GP, Rolston KV (2000) Zygomycosis in the 1990s in a tertiary-care cancer center. Clin Infect Dis 30:851–856 Kontoyiannis DP, Lionakis MS, Lewis RE, Chamilos G, Healy M, Perego C, Safdar A, Kantarjian H, Champlin R, Walsh TJ, Raad II (2005) Zygomycosis in a tertiary-care cancer center in the era of Aspergillus-active antifungal therapy: a case-control observational study of 27 recent cases. J Infect Dis 191:1350–1360 Kurtzman CP, Fell JW (1998) The yeasts, a taxonomic study, 4th edn. Elsevier Science BV, Amsterdam, pp 90–1000 Kurtzman CP, Robnett CJ (1991) Phylogenetic relationships among species of Saccharomyces, Schizosaccharomyces, Debaryomyces, and Schwanniomyces determined from partial ribosomal RNA sequences. Yeast 7:61–72 Kurtzman CP, Robnett CJ (1994) Synonymy of the yeast genera Wingea and Debaryomyces. Antonie Van Leeuwenhoek 66:337–342 Kurtzman CP, Robnett CJ (1998) Identification and phylogeny of ascomycetous yeasts from analysis of nuclear large subunit (26S) ribosomal DNA partial sequences. Antonie Van Leeuwenhoek 73:331–371 Kwon-Chung KJ, Boekhout T, Fell JW (2002) Proposal to conserve the name Cryptococcus gattii against C hondurianus and C bacillisporus. Taxon 51:804–806 Latgé JP (1999) Aspergillus fumigatus and aspergillosis. Clin Microbiol Rev 12:310–350 Latgé JP, Kobayashi H, Debeaupuis JP, Diaquin M, Sarfati J, Wieruszeski JM, Parra E, Bouchara JP, Fournet B (1994) Chemical and immunological characterization of the extracellular galactomannan of Aspergillus fumigatus. Infect Immun 62:5424–5433 Lau A, Chen S, Sorrell T, Carter D, Malik R, Martin P, Halliday C (2007) Development and clinical application of a panfungal PCR assay to detect and identify fungal DNA in tissue specimens. J Clin Microbiol 45:380–385 Lemieux C, St-Germain G, Vincelette J, Kaufman L, de Repentigny L (1990) Collaborative evaluation of antigen detection by a commercial latex agglutination test and enzyme immunoassay in the diagnosis of invasive candidiasis. J Clin Microbiol 28:249–253 Li YL, Leaw SN, Chen JH, Chang HC, Chang TC (2003) Rapid identification of yeasts commonly found in positive blood cultures byamplification of the internal transcribed spacer regions 1 and 2. Eur J Clin Microbiol Infect Dis 22:693–696 17 DNA-Based Detection of Human Pathogenic Fungi 407 Lindsley MD, Hurst SF, Iqbal NJ, Morrison CJ (2001) Rapid identification of dimorphic and yeastlike fungal pathogens using specific DNA probes. J Clin Microbiol 39:3505–3511 Linton CJ, Borman AM, Cheung G, Holmes AD, Szekely A, Palmer MD, Bridge PD, Campbell CK, Johnson EM (2007) Molecular identification of unusual pathogenic yeast isolates by large ribosomal subunit gene sequencing: 2 years of experience at the United kingdom mycology reference laboratory. J Clin Microbiol 45:1152–1158 Lionakis MS, Kontoyiannis DP (2004) Fusarium infections in critically ill patients. Semin Respir Crit Care Med 25:159–169 Liu D, Pearce L, Lilley G, Coloe S, Baird R, Pedersen J (2002) PCR identification of dermatophyte fungi Trichophyton rubrum, T soudanense and T gourvilii. J Med Microbiol 51:117–122 Löffler J, Hebart H, Schumacher U, Reitze H, Einsele H (1997) Comparison of different methods for extraction of DNA of fungal pathogens from cultures and blood. J Clin Microbiol 35:3311–3312 Löffler J, Hebart H, Sepe S, Schumcher U, Klingebiel T, Einsele H (1998) Detection of PCRamplified fungal DNA by using a PCR-ELISA system. Med Mycol 36:275–279 Löffler J, Hebart H, Bialek R, Hagmeyer L, Schmidt D, Serey FP, Hartmann M, Eucker J, Einsele H (1999) Contaminations occurring in fungal PCR assays. J Clin Microbiol 37:1200–1202 Löffler J, Hebart H, Brauchle U, Schumacher U, Einsele H (2000) Comparison between plasma and whole blood specimens for detection of Aspergillus DNA by PCR. J Clin Microbiol 38:3830–3833 Logotheti M, Kotsovili-Tseleni A, Arsenis G, Legakis NI (2009) Multiplex PCR for the discrimination of A fumigatus, A flavus, A niger and A terreus. J Microbiol Methods 76:209–211 Lott TJ, Burns BM, Zancope-Oliveira R, Elie CM, Reiss E (1998) Sequence analysis of the internal transcribed spacer 2 (ITS2) from yeast species within the genus Candida. Curr Microbiol 36:63–69 Maaroufi Y, Heymans C, De Bruyne JM, Duchateau V, Rodriguez-Villalobos H, Aoun M, Crokaert F (2003) Rapid detection of Candida albicans in clinical blood samples by using a TaqMan-based PCR assay. J Clin Microbiol 41:3293–3298 Machouart M, Larche J, Burton K, Collomb J, Maurer P, Cintrat A, Biava MF, Greciano S, Kuijpers AF, Contet-Audonneau N, de Hoog GS, Gerard A, Fortier B (2006) Genetic identification of the main opportunistic Mucorales by PCR-restriction fragment length polymorphism. J Clin Microbiol 44:805–810 Makimura K, Murayama SY, Yamaguchi H (1994) Specific detection of Aspergillus and Penicillium species from respiratory specimens by polymerase chain reaction (PCR). Jpn J Med Sci Biol 47:141–156 Makimura K, Tamura Y, Mochizuki T, Hasegawa A, Tajiri Y, Hanazawa R, Uchida K, Saito H, Yamaguchi H (1999) Phylogenetic classification and species identification of dermatophyte strains based on DNA sequences of nuclear ribosomal internal transcribed spacer 1 regions. J Clin Microbiol 37:920–924 Mancini N, Clerici D, Diotti R, Perotti M, Ghidoli N, De Marco D, Pizzorno B, Emrich T, Burioni R, Ciceri F, Clementi M (2008) Molecular diagnosis of sepsis in neutropenic patients with haematological malignancies. J Med Microbiol 57:601–604 Martagon-Villamil J, Shrestha N, Sholtis M, Isada CM, Hall GS, Bryne T, Lodge BA, Reller LB, Procop GW (2003) Identification of Histoplasma capsulatum from culture extracts by real-time PCR. J Clin Microbiol 41:1295–1298 Martin C, Roberts D, van Der Weide M, Rossau R, Jannes G, Smith T, Maher M (2000) Development of a PCR-based line probe assay for identification of fungal pathogens. J Clin Microbiol 38:3735–3742 McCullough MJ, DiSalvo AF, Clemons KV, Park P, Stevens DA (2000) Molecular epidemiology of Blastomyces dermatitidis. Clin Infect Dis 30:328–335 Medoff G, Dismukes WE, Pappagianis D, Diamond R, Gallis HA, Drutz D (1992) Evaluation of new antifungal drugs for the treatment of systemic fungal infections Infectious Diseases Society of America and the Food and Drug Administration. Clin Infect Dis 15:S274–S281 408 L. Putignani et al. Mekha N, Sugita T, Ikeda R, Nishikawa A, Poonwan N (2007) Real-time PCR assay to detect DNA in sera for the diagnosis of deep-seated trichosporonosis. Microbiol Immunol 51:633–635 Melcher GP, Reed KD, Rinaldi MG, Lee JW, Pizzo PA, Walsh TJ (1991) Demonstration of a cell wall antigen cross-reacting with cryptococcal polysaccharide in experimental disseminated trichosporonosis. J Clin Microbiol 29:192–196 Melchers WJ, Verweij PE, van den Hurk P, van Belkum A, De Pauw BE, Hoogkamp-Korstanje JA, Meis JF (1994) General primer-mediated PCR for detection of Aspergillus species. J Clin Microbiol 32:1710–1717 Mengoli C, Cruciani M, Barnes RA, Loeffler J, Donnelly JP (2009) Use of PCR for diagnosis of invasive aspergillosis: systematic review and meta-analysis. Lancet Infect Dis 9:89–96 Mennink-Kersten MA, Klont RR, Warris A, Op den Camp HJ, Verweij PE (2004) Bifidobacterium lipoteichoic acid and false ELISA reactivity in aspergillus antigen detection. Lancet 363:325–327 Metwally L, Hogg G, Coyle PV, Hay RJ, Hedderwick S, McCloskey B, O’Neill HJ, Ong GM, Thompson G, Webb CH, McMullan R (2007) Rapid differentiation between fluconazolesensitive and -resistant species of Candida directly from positive blood-culture bottles by real-time PCR. J Med Microbiol 56:964–970 Metwally L, Fairley DJ, Coyle PV, Hay RJ, Hedderwick S, McCloskey B, O’Neill HJ, Webb CH, Elbaz W, McMullan R (2008a) Improving molecular detection of Candida DNA in whole blood: comparison of seven fungal DNA extraction protocols using real-time PCR. J Med Microbiol 57:296–303 Metwally L, Fairley DJ, Coyle PV, Hay RJ, Hedderwick S, McCloskey B, O’Neill HJ, Webb CH, McMullan R (2008b) Comparison of serum and whole-blood specimens for the detection of Candida DNA in critically ill, non-neutropenic patients. J Med Microbiol 57:1269–1272 Meyer W, Castañeda A, Jackson S, Huynh M, Castañeda E, IberoAmerican Cryptococcal Study Group (2003) Molecular typing of IberoAmerican Cryptococcus neoformans isolates. Emerg Infect Dis 9:189–195 Milam CP, Fenske NA (1989) Chromoblastomycosis. Dermatol Clin 7:219–225 Millon L, Barale T, Julliot MC, Martinez J, Mantion G (1995) Interference by hydroxyethyl starch used for vascular filling in latex agglutination test for cryptococcal antigen. J Clin Microbiol 33:1917–1919 Mishra PK, Fox RT, Culham A (2003) Development of a PCR-based assay for rapid and reliable identification of pathogenic Fusaria. FEMS Microbiol Lett 218:329–332 Mitchell TG, Perfect JR (1995) Cryptococcosis in the era of AIDS–100 years after the discovery of Cryptococcus neoformans. Clin Microbiol Rev 8:515–548 Mitchell TG, Freedman EZ, White TJ, Taylor JW (1994) Unique oligonucleotide primers in PCR for identification of Cryptococcus neoformans. J Clin Microbiol 32:253–255 Miyagi H, Yamamoto Y, Kanamori S, Taira K, Asato Y, Myint CK, Kayo S, Hosokawa A, Hagiwara K, Uezato H (2008) Case of chromoblastomycosis appearing in an Okinawa patient with a medical history of Hansen’s disease. J Dermatol 35:354–3561 Miyakawa Y, Mabuchi T, Kagaya K, Fukazawa Y (1992) Isolation and characterization of a species specific DNA fragment for detection of Candida albicans by polymerase chain reaction. J Clin Microbiol 30:894–900 Miyakawa Y, Mabuchi T, Fukazawa Y (1993) New method for detection of Candida albicans in human blood by polymerase chain reaction. J Clin Microbiol 31:3344–3347 Miyazaki T, Kohno S, Mitsutake K, Maesaki S, Tanaka K, Hara K (1995a) (1 ! 3)-b-D-Glucan in culture fluid of fungi activates factor G A limulus coagulation factor. J Clin Lab Anal 9:334–339 Miyazaki T, Kohno S, Mitsutake K, Maesaki S, Tanaka KI, Ishikawa N, Hara K (1995b) Plasma (1–3)-b-D-glucan and fungal antigenemia in patients with candidemia, aspergillosis, and cryptococcosis. J Clin Microbiol 33:3115–3118 Mochizuki T, Tanabe H, Kawasaki M, Ishizaki H, Jackson CJ (2003) Rapid identification of Trichophyton tonsurans by PCR-RFLP analysis of ribosomal DNA regions. J Dermatol Sci 32:25–32 17 DNA-Based Detection of Human Pathogenic Fungi 409 Morace G, Sanguinetti M, Posteraro B, Lo Cascio G, Fadda G (1997) Identification of various medically important Candida species in clinical specimens by PCR-restriction enzyme analysis. J Clin Microbiol 35:667–672 Morace G, Pagano L, Sanguinetti M, Posteraro B, Mele L, Equitani F, D’Amore G, Leone G, Fadda G (1999) PCR-restriction enzyme analysis for detection of Candida DNA in blood from febrile patients with hematological malignancies. J Clin Microbiol 37:1871–1875 Mounts A, Deepe GS (1998) Simultaneous infection with Blastomyces dermatitidis and Cryptococcus neoformans. Med Mycol 36:47–50 Mulè G, González-Jaén MT, Hornok L, Nicholson P, Waalwijk C (2005) Advances in molecular diagnosis of toxigenic Fusarium species: a review. Food Addit Contam 22:316–323 Müller E, von Arx JA (1973) Pyrenomycetes: Meliolales, Coronophorales, Sphaeriales. In: Ainsworth GC, Sparrow FK, Sussman AS (eds) The fungi: an advanced treatise, vol 4A. Academic Press, New York, NY, pp 87–132 Müller FM, Werner KE, Kasai M, Francesconi A, Chanock SJ, Walsh TJ (1998) Rapid extraction of genomic DNA from medically important yeasts and filamentous fungi by high-speed cell disruption. J Clin Microbiol 36:1625–2629 Nagao K, Ota T, Tanikawa A, Takae Y, Mori T, Udagawa S, Nishikawa T (2005) Genetic identification and detection of human pathogenic Rhizopus species, a major mucormycosis agent, by multiplex PCR based on internal transcribed spacer region of rRNA gene. J Dermatol Sci 39:23–31 Naidu PS, Zhang YZ, Reddy CA (1990) Characterization of a new lignin peroxidase gene (GLC6) from Phanerochaete chrysosporium. Biochem Biophys Res Commun 173:994–1000 Navarro EE, Almario JS, Schaufele RL, Bacher J, Walsh TJ (1997) Quantitative urine cultures do not reliably detect renal candidiasis in rabbits. J Clin Microbiol 35:3292–3297 Nucci M, Marr KA, Queiroz-Telles F, Martins CA, Trabasso P, Costa S, Voltarelli JC, Colombo AL, Imhof A, Pasquini R, Maiolino A, Souza CA, Anaissie E (2004) Fusarium infection in hematopoietic stem cell transplant recipients. Clin Infect Dis 38:1237–1242 Ogawa H, Summerbell RC, Clemons KV, Koga T, Ran YP, Rashid A, Sohnle PG, Stevens DA, Tsuboi R (1998) Dermatophytes and host defence in cutaneous mycoses. Med Mycol 36:166–173 Ostrosky-Zeichner L (2003) New approaches to the risk of Candida in the intensive care unit. Curr Opin Infect Dis 16:533–537 Pagano L, Fianchi L, Caramatti C, D’Antonio D, Melillo L, Caira M, Masini L, Todeschini G, Girmenia C, Martino B, Cinieri S, Martino P, Del Bavero A, Gruppo Italiano Malattie EMatologiche dell’Adulto Infection Program (2004) Cryptococcosis in patients with hematologic malignancies A report from GIMEMA-infection. Haematologica 89:852–856 Pappagianis D (2001) Serologic studies in coccidioidomycosis. Semin Respir Infect 16:242–250 Pappas PG (2004) Blastomycosis. Semin Respir Crit Care Med 25:1131–21 Paschoal RC, Hirata MH, Hirata RC, Melhem Mde S, Dias AL, Paula CR (2004) Neurocryptococcosis: diagnosis by PCR method. Rev Inst Med Trop Sao Paulo 46:203–207 Patterson JE (2005) In: Mandell GL, Bennet J, Dolin R (eds) Principles and practice of infectious diseases, vol 2. Elsevier Churchill Livingstone, New York, NY, pp 2958–2973 Patterson JE, Peters J, Calhoon JH, Levine S, Anzueto A, Al-Abdely H, Sanchez R, Patterson TF, Rech M, Jorgensen JH, Rinaldi MG, Sako E, Johnson S, Speeg V, Halff GA, Trinkle JK (2000) Investigation and control of aspergillosis and other filamentous fungal infections in solid organ transplant recipients. Transpl Infect Dis 2:22–28 Perfect JR, Casadevall A (2002) Cryptococcosis. Infect Dis Clin North Am 16:837–874 Peterson SW, Kurtzman CP (1991) Ribosomal RNA sequence divergence among sibling species of yeasts. Syst Appl Microbiol 14:124–129 Pfaller MA, Diekema DJ (2004) International fungal surveillance participant group 12 years of fluconazole in clinical practice: global trends in species distribution and fluconazole susceptibility of bloodstream isolates of Candida. Clin Microbiol Infect 10(Suppl 1):11–23 Pfaller MA, Diekema DJ (2007) Epidemiology of invasive candidiasis: a persistent public health problem. Clin Microbiol Rev 20:133–163 410 L. Putignani et al. Pham AS, Tarrand JJ, May GS, Lee MS, Kontoyiannis DP, Han XY (2003) Diagnosis of invasive mold infection by real-time quantitative PCR. Am J Clin Pathol 119:38–44 Phillips P, Dowd A, Jewesson P, Radigan G, Tweeddale MG, Clarke A, Greere I, Kelly M (1990) Nonvalue of antigen detection immunoassays for diagnosis of candidemia. J Clin Microbiol 28:2320–2326 Piepenbring M, Cáceres Mendez OA, Espino Espinoza AA, Kirschner R, Schöfer H (2007) Chromoblastomycosis caused by Chaetomium funicola: a case report from Western Panama. Br J Dermatol 157:1025–1029 Playford EG, Kong F, Sun Y, Wang H, Halliday C, Sorrell TC (2006) Simultaneous detection and identification of Candida, Aspergillus, and Cryptococcus species by reverse line blot hybridization. J Clin Microbiol 44:876–880 Polesky A, Kirsch CM, Snyder LS, LoBue P, Kagawa FT, Dykstra BJ, Wehner JH, Catanzaro A, Ampel NM, Stevens DA (1999) Airway coccidioidomycosis – report of cases and review. Clin Infect Dis 28:1273–1280 Pounder JI, Hansen D, Woods GL (2006) Identification of Histoplasma capsulatum, Blastomyces dermatitidis, and Coccidioides species by repetitive-sequence-based PCR. J Clin Microbiol 44:2977–2982 Powell BL, Drutz DJ, Huppert M, Sun SH (1983) Relationship of progesterone- and estradiolbinding proteins in Coccidioides immitis to coccidioidal dissemination in pregnancy. Infect Immun 40:478–485 Prariyachatigul C, Chaiprasert A, Meevootisom V, Pattanakitsakul S (1996) Assessment of a PCR technique for the detection and identification of Cryptococcus neoformans. J Med Vet Mycol 34:251–258 Prentice HG, Kibbler CC, Prentice AG (2000) Towards a targeted, risk-based, antifungal strategy in neutropenic patients. Br J Haematol 110:273–284 Prevost E, Bannister E (1981) Detection of yeast septicemia by biphasic and radiometric methods. J Clin Microbiol 12:655–660 Prince HE, Yeh C, Alem N, Asalkhou M, Hamedi N, Alem N, Alem M (2008) Evaluation of enzyme-linked immunosorbent assays for detecting circulating antibodies to Candida albicans. J Clin Lab Anal 22:234–238 Putignani L, Antonucci G, Paglia MG, Vincenti L, Festa A, De Mori P, Loiacono L, Visca P (2008a) Cryptococcal lymphadenitis as a manifestation of immune reconstitution inflammatory syndrome in an HIV-positive patient: a case report and review of the literature. Int J Immunopathol Pharmacol 21:751–756 Putignani L, Paglia MG, Bordi E, Nebuloso E, Pucillo LP, Visca P (2008b) Identification of clinically relevant yeast species by DNA sequence analysis of the D2 variable region of the 25–28S rRNA gene. Mycoses 51:209–227 Raad I, Hanna H, Huaringa A, Sumoza D, Hachem R, Albitar M (2002) Diagnosis of invasive pulmonary aspergillosis using polymerase chain reaction-based detection of aspergillus in BAL. Chest 121:1171–1176 Radford A (1993) A fungal phylogeny based upon orotidine-50 -monophosphate decarboxylase. J Mol Evol 36:389–395 Raimondi A, Ticozzi R, Sala G, Bellotti MG (2007) Genotype-based differentiation of the Cryptococcus neoformans serotypes by combined PCR-RFLP analysis of the capsuleassociated genes CAP10 and CAP59. Med Mycol 45:491–501 Ralph ED, Hussain Z (1996) Chronic meningitis caused by Candida albicans in a liver transplant recipient: usefulness of the polymerase chain reaction for diagnosis and for monitoring treatment. Clin Infect Dis 23:191–192 Raman C, Khardori N, Von Behren LA, Wheat LJ, Tewary RP (1990) Evaluation of an ELISA for the detection of anti-Histoplasma ribosomal and antihistoplasmin antibodies in histoplasmosis. J Clin Lab Anal 4:199–207 Rantakokko-Jalava K, Laaksonen S, Issakainen J, Vauras J, Nikoskelainen J, Viljanen MK, Salonen J (2003) Semiquantitative detection by real-time PCR of Aspergillus fumigatus in 17 DNA-Based Detection of Human Pathogenic Fungi 411 bronchoalveolar lavage fluids and tissue biopsy specimens from patients with invasive aspergillosis. J ClinMicrobiol 41:4304–4311 Rappelli P, Are R, Casu G, Fiori PL, Cappuccinelli P, Aceti A (1998) Development of a nested PCR for detection of Cryptococcus neoformans in cerebrospinal fluid. J Clin Microbiol 36:3438–3440 Rath PM, Ansorg R (2000) Identification of medically important Aspergillus species by single strand conformational polymorphism (SSCP) of the PCR-amplified intergenic spacer region. Mycoses 43:381–386 Reddy LV, Kumar A, Kurup VP (1993) Specific amplification of Aspergillus fumigatus DNA by polymerase chain reaction. Mol Cell Probes 7:121–126 Reiss E, Morrison CJ (1993) Nonculture methods for diagnosis of disseminated candidiasis. Clin Microbiol 6:311–323 Revankar SG (2006) Phaeohyphomycosis. Infect Dis Clin North Am 20:609–620 Revankar SG, Patterson JE, Sutton DA, Pullen R, Rinaldi MG (2002) Disseminated phaeohyphomycosis: review of an emerging mycosis. Clin Infect Dis 34:467–476 Revankar SG, Sutton DA, Rinaldi MG (2004) Primary central nervous system phaeohyphomycosis: a review of 101 cases. Clin Infect Dis 38:206–216 Ribes JA, Vanover-Sams CL, Baker DJ (2000) Zygomycetes in human disease. Clin Microbiol Rev 13:236–301 Rickerts V, Loeffler J, Bohme A, Einsele H, Just-Nubling G (2001) Diagnosis of disseminated zygomycosis using a polymerase chain reaction assay. Eur J Clin Microbiol Infect Dis 20:744–745 Rickerts V, Bialek R, Tintelnot K, Jacobi V, Just-Nübling G (2002) Rapid PCR-based diagnosis of disseminated histoplasmosis in an AIDS patient. Eur J Clin Microbiol Infect Dis 21:821–823 Rickerts V, Just-Nübling G, Konrad F, Kern J, Lambrecht E, Böhme A, Jacobi V, Bialek R (2006) Diagnosis of invasive aspergillosis and mucormycosis in immunocompromised patients by seminested PCR assay of tissue samples. Eur J Clin Microbiol Infect Dis 25:8–13 Roden MM, Zaoutis TE, Buchanan WL, Knudsen TA, Sarkisova TA, Schaufele RL, Sein M, Sein T, Chiou CC, Chu JH, Kontoyiannis DP, Walsh TJ (2005) Epidemiology and outcome of zygomycosis: a review of 929 reported cases. Clin Infect Dis 41:634–653 Rodriguez LJ, Rex JH, Anaissie EJ (1996) Update on invasive candidiasis. Adv Pharmacol 137:349–400 Rodriguez LJ, Anaissie EJ, Rex JH (2000) Pneumonia due to Candida species. In: Sarosi GA, Davies SF (eds) Fungal disease of the lung, 3rd edn. Lippincott Williams & Wilkins, Philadelphia, pp 115–122 Rodriguez-Tudela J, Diaz-Guerra T, Mellado E, Cano V, Tapia C, Perkins A, Gomez-Lopez A, Rodero L, Cuenca-Estrealla M (2005) Susceptibility patterns and molecular identification of Trichosporon species Antimicrob. Agents Chemother 49:4026–4034 Rüping MJ, Vehreschild JJ, Cornely OA (2008) Patients at high risk of invasive fungal infections: when and how to treat. Drugs 68:1941–1962 Saha DC, Xess I, Jain N (2008) Evaluation of conventional and serological methods for rapid diagnosis of cryptococcosis. Indian J Med Res 12:483–488 Saiman L, Ludington E, Pfaller M, Rangel-Frausto S, Wiblin RT, Dawson J, Blumberg HM, Patterson JE, Rinaldi M, Edwards JE, Wenzel RP, Jarvis W (2000) Risk factors for candidemia in neonatal intensive care unit patients. Pediat Inf Dis J 19:319–324 Sanche SE, Sutton DA, Rinaldi MG (2003) Dematiaceous fungi. In: Anaissie EJ, McGinnis MR, Pfaller MA (eds) Clinical mycology. Churchill Livingstone, Philadelphia Sandhu GS, Kline BC, Stockman L, Roberts GD (1995) Molecular probes for diagnosis of fungal infections. J Clin Microbiol 33:2913–2919 Sanguinetti M, Posteraro B, Pagano L, Pagliari G, Fianchi L, Mele L, La Sorda M, Franco A, Fadda G (2003) Comparison of real-time PCR, conventional PCR, and galactomannan antigen detection by enzyme-linked immunosorbent assay using bronchoalveolar lavage fluid samples from hematology patients for diagnosis of invasive pulmonary aspergillosis. J Clin Microbiol 41:3922–3925 412 L. Putignani et al. Sanguinetti M, Porta R, Sali M, La Sorda M, Pecorini G, Fadda G, Posteraro B (2007) Evaluation of VITEK 2 and RapID yeast plus systems for yeast species identification: experience at a large clinical microbiology laboratory. J Clin Microbiol 45:1343–1346 Sarkar G, Sommer S (1990) Shedding light on PCR contamination. Nature 347:340–341 Schabereiter-Gurtner C, Selitsch B, Rotter ML, Hirschl AM, Willinger B (2007) Development of novel real-time PCR assays for detection and differentiation of eleven medically important Aspergillus and Candida species in clinical specimens. J Clin Microbiol 45:906–914 Schwarz P, Bretagne S, Gantier JC, Garcia-Hermoso D, Lortholary O, Dromer F, Dannaoui E (2006) Molecular identification of zygomycetes from culture and experimentally infected tissues. J Clin Microbiol 44:340–349 Segal BH, Walsh TJ (2006) Current approaches to diagnosis and treatment of invasive aspergillosis. Am J Respir Crit Care Med 173:707–717 Selvarangan R, Bui U, Limaye AP, Cookson BT (2003) Rapid identification of commonly encountered Candida species directly from blood culture bottles. J Clin Microbiol 41:5660–5664 Sendid B, Tabouret M, Poirot JL, Mathieu D, Fruit J, Poulain D (1999) New enzyme immunoassay for sensitive detection of circulating Candida albicans mannan and antimannan antibodies: useful combined test for diagnosis of systemic candidiasis. J Clin Microbiol 37:1510–1517 Sheppard DC, Locas MC, Restieri C, Laverdiere M (2008) Utility of the germ tube test for direct identification of Candida albicans from positive blood culture bottles. J Clin Microbiol 46:3508–3509 Singh N, Lortholary O, Alexander BD, Gupta KL, John GT, Pursell K, Munoz P, Klintmalm GB, Stosor V, del Busto R, Limaye AP, Somani J, Lyon M, Houston S, House AA, Pruett TL, Orloff S, Humar A, Dowdy L, Garcia-Diaz J, Kalil AC, Fisher RA, Husain S, Cryptococcal Collaborative Transplant Study Group (2005) An immune reconstitution syndrome-like illness associated with Cryptococcus neoformans infection in organ transplant recipients. Clin Infect Dis 40:1756–1761 Sivakumar VG, Shankar P, Nalina K, Menon T (2009) Use of CHROMagar in the differentiation of common species of Candida. Mycopathologia 167:47–49 Skladny H, Buchheidt D, Baust C, Krieg-Schneider F, Seifarth W, Leib-Mösch C, Hehlmann R (1999) Specific detection of Aspergillus species in blood and bronchoalveolar lavage samples of immunocompromised patients by two-step PCR. J Clin Microbiol 37:3865–3871 Snow RM, Dismukes WE (1975) Cryptococcal meningitis: diagnostic value of cryptococcal antigen in cerebrospinal fluid. Arch Intern Med 135:1155–1157 Spiess B, Buchheidt D, Baust C, Skladny H, Seifarth W, Zeilfelder U, Leib-Mösch C, Mörz H, Hehlmann R (2003) Development of a LightCycler PCR assay for detection and quantification of Aspergillus fumigatus DNA in clinical samples from neutropenic patients. J Clin Microbiol 41:1811–1818 Spiess B, Seifarth W, Hummel M, Frank O, Fabarius A, Zheng C, Mörz H, Hehlmann R, Buchheidt D (2007) DNA microarray-based detection and identification of fungal pathogens in clinical samples from neutropenic patients. J Clin Microbiol 45:3743–3753 Spreadbury C, Holden D, Aufauvre-Brown A, Bainbridge B, Cohen J (1993) Detection of Aspergillus fumigatus by polymerase chain reaction. J Clin Microbiol 31:615–621 Stockman L, Clark KA, Hunt JM, Roberts GD (1993) Evaluation of commercially available acridinium ester-labeled chemiluminescent DNA probes for culture identification of Blastomyces dermatitidis, Coccidioides immitis, Cryptococcus neoformans, and Histoplasma capsulatum. J Clin Microbiol 31:845–850 Suarez F, Lortholary O, Buland S, Rubio MT, Ghez D, Mahé V, Quesne G, Poirée S, Buzyn A, Varet B, Berche P, Bougnoux ME (2008) Detection of circulating Aspergillus fumigatus DNA by real-time PCR assay of large serum volumes improves early diagnosis of invasive aspergillosis in high-risk adult patients under hematologic surveillance. J Clin Microbiol 46:3772–3777 Sugita T, Nishikawa A, Shinoda T (1998) Rapid detection of species of the opportunistic yeast Trichosporon by PCR. J Clin Microbiol 36:1458–1460 17 DNA-Based Detection of Human Pathogenic Fungi 413 Sugita TA, Nishikawa A, Ikeda R, Shinoda T (1999) Identification of medically relevant Trichosporon species based on sequences of internal transcribed spacer regions and construction of a database for Trichosporon identification. J Clin Microbiol 37:1985–1993 Sugita T, Nakajima M, Ikeda R, Niki Y, Matsushima T, Shinoda T (2001) A nested PCR assay to detect DNA in sera for the diagnosis of deep-seated trichosporonosis. Microbiol Immunol 45:143–148 Sullivan DJ, Coleman DC (1998) Candida dubliniensis; characteristics and identification. J Clin Microbiol 31:229–334 Swanink CM, Meis JF, Rijs AJ, Donnelly JP, Verweij PE (1997) Specificity of a sandwich enzyme-linked immunosorbent assay for detecting Aspergillus galactomannan. J Clin Microbiol 35:257–260 Takahashi T, Goto M, Kanda T, Iwamoto A (2003) Utility of testing bronchoalveolar lavage fluid for cryptococcal ribosomal DNA. J Int Med Res 31:324–329 Tanaka KI, Miyazaki T, Maesaki S, Mitsutake K, Kakeya H, Yamamoto Y, Yanagihara K, Hossain MA, Tashiro T, Kohno S (1996) Detection of Cryptococcus neoformans gene in patients with pulmonary cryptococcosis. J Clin Microbiol 34:2826–2828 Tang CM, Holden DW, Aufauvre-Brown A, Cohen J (1993) The detection of Aspergillus spp by the polymerase chain reaction and its evaluation in bronchoalveolar lavage fluid. Am Rev Respir Dis 148:1313–1317 Tavanti A, Davidson AD, Gow NA, Maiden MC, Odds FC (2005) Candida orthopsilosis and Candida metapsilosis spp nov to replace Candida parapsilosis groups II and III. J Clin Microbiol 43:284–292 Torres HA, Raad II, Kontoyiannis DP (2003) Infections caused by Fusarium species. J Chemother 15:28–35 Tortorano AM, Peman J, Bernhardt H, Klingspor L, Kibbler CC, Faure O, Biraghi E, Canton E, Zimmermann K, Seaton S, Grillot R (2004) Epidemiology of candidaemia in Europe: results of 28-month European Confederation of Medical Mycology (ECMM) hospital-based surveillance study. Eur J Clin Microbiol Infect Dis 23:317–322 Trofa D, Gácser A, Nosanchuk JD (2008) Candida parapsilosis, an emerging fungal pathogen. Clin Microbiol Rev 21:606–625 Tsai HF, Liu JS, Staben C, Christensen MJ, Latch GCM, Siegel MR, Schardl CL (1994) Evolutionary diversification of fungal endophytes of tall fescue grass by hybridization with Epichloë species. Proc Natl Acad Sci USA 91:2542–2546 Turenne CY, Sanche SE, Hoban DJ, Karlowsky JA, Kabani AM (1999) Rapid identification of fungi by using the ITS2 genetic region and an automated fluorescent capillary electrophoresis system. J Clin Microbiol 37:1846–1851 Turin L, Riva F, Galbiati G, Cainelli T (2000) Fast, simple and highly sensitive double-rounded polymerase chain reaction assay to detect medically relevant fungi in dermatological specimens. Eur J Clin Invest 30:511–518 Umeyama T, Sano A, Kamei K, Niimi M, Nishimura K, Uehara Y (2006) Novel approach to designing primers for identification and distinction of the human pathogenic fungi Coccidioides immitis and Coccidioides posadasii by PCR amplification. J Clin Microbiol 44:1859–1862 Van Burik JA, Myerson D, Schreckhise RW, Bowden RA (1998) Panfungal PCR assay for detection of fungal infection in human blood specimens. J Clin Microbiol 36:1169–1175 van de Peer Y, Chapelle S, De Wachter R (1996) A quantitative map of nucleotide substitution rates in bacterial rRNA. Nucleic Acids Res 24:3381–3391 van de Peer Y, Jansen J, De Rijk J, De Wachter R (1997) Database on the structure of small ribosomal subunit RNA. Nucleic Acids Res 25:111–116 Veer P, Patwardhan NS, Damle AS (2007) Study of onychomycosis: prevailing fungi and pattern of infection. Indian J Med Microbiol 25:53–56 Vidal MS, Castro LG, Cavalcante SC, Lacaz CS (2004) Highly specific and sensitive, immunoblot-detected 54 kDa antigen from Fonsecaea pedrosoi. Med Mycol 42:511–515 Vilchez RA, Fung J, Kusne S (2002) Cryptococcosis in organ transplant recipients: an overview. Am J Transplant 2:575–580 414 L. Putignani et al. Vollmer T, Störmer M, Kleesiek K, Dreier J (2008) Evaluation of novel broad-range real-time PCR assay for rapid detection of human pathogenic fungi in various clinical specimens. J Clin Microbiol 46:1919–1926 Wadlin JK, Hanko G, Stewart R, Pape J, Nachamkin I (1999) Comparison of three commercial systems for identification of yeasts commonly isolated in the clinical microbiology laboratory. J Clin Microbiol 37:1967–1970 Wahyuningsih R, Freisleben HJ, Sonntag HG, Schnitzler P (2000) Simple and rapid detection of Candida albicans DNA in serum by PCR for diagnosis of invasive candidiasis. J Clin Microbiol 38:3016–3021 Walker WF, Doolittle WF (1982) Redividing the basidiomycetes on the basis of 5S rRNA sequences. Nature 299:723–724 Walsh TJ, Francesconi A, Kasai M, Chanock SJ (1995) PCR and single-strand conformational polymorphism for recognition of medically important opportunistic fungi. J Clin Microbiol 33:3216–3220 Wang L, Yokoyama K, Miyaji M, Nishimura K (2000) Mitochondrial cytochrome b gene analysis of Aspergillus fumigatus and related species. J Clin Microbiol 38:1352–1358 Wang SM, Hsu CH, Chang JH (2008) Congenital candidiasis. Pediatr Neonatol 49:94–96 Wheat JL (2003) Current diagnosis of histoplasmosis. Trends Microbiol 11:488–494 Wheat LJ, Kauffman CA (2003) Histoplasmosis. Infect Dis Clin North Am 17:1–19 Wheat LJ, Connolly-Stringfield PA, Williams B, Connolly K, Blair R, Bartlett M (1992) Diagnosis of histoplasmosis in patients with the acquired immunodeficiency syndrome by detection of Histoplasma capsulatum polysaccharide antigen in bronchoalveolar lavage fluid. Am Rev Respir Dis 145:1421–1424 Wheat LJ, Carringer T, Brizendine E, Connolly P (2002) Diagnosis of histoplasmosis by antigen detection based upon experience at the histoplasmosis reference laboratory. Diagn Microbiol Infect Dis 43:29–37 White PL, Shetty A, Barnes RA (2003) Detection of seven Candida species using the Light-Cycler system. J Med Microbiol 52:229–238 White PL, Williams DW, Kuriyama T, Samad SA, Lewis MA, Barnes RA (2004) Detection of Candida in concentrated oral rinse cultures by real-time PCR. J Clin Microbiol 42:2101–2107 White PL, Barton R, Guiver M, Linton CJ, Wilson S, Smith M, Gomez BL, Carr MJ, Kimmitt PT, Seaton S, Rajakumar K, Holyoake T, Kibbler CC, Johnson E, Hobson RP, Jones B, Barnes RA (2006) A consensus on fungal polymerase chain reaction diagnosis?: a United KingdomIreland evaluation of polymerase chain reaction methods for detection of systemic fungal infections. J Mol Diagn 8:376–384 Whittier S, Hopfer RL, Gilligan P (1994) Elimination of false-positive serum reactivity in latex agglutination test for cryptococcal antigen in human immunodeficiency virus-infected population. J Clin Microbiol 32:2158–2161 Williamson EC, Leeming JP (1999) Molecular approaches for the diagnosis and epidemiological investigation of Aspergillus infection. Mycoses 42:S7–10 Williamson EC, Leeming JP, Palmer HM, Steward CG, Warnock D, Marks DI, Millar MR (2000) Diagnosis of invasive aspergillosis in bone marrow transplant recipients by polymerase chain reaction. Br J Haematol 108:132–139 Wise MG, Healy M, Reece K, Smith R, Walton D, Dutch W, Renwick A, Huong J, Young S, Tarrand J, Kontoyiannis DP (2007) Species identification and strain differentiation of clinical Candida isolates using the DiversiLab system of automated repetitive sequence-based PCR. J Med Microbiol 56:778–787 Yamada Y, Makimura K, Uchida K, Yamaguchi H, Osumi M (2004) Phylogenetic relationships among medically important yeasts based on sequences of mitochondrial large subunit ribosomal RNA gene. Mycoses 47:24–28 Yeo SF, Wong B (2002) Current status of nonculture methods for diagnosis of invasive fungal infections. Clin Microbiol Rev 15:465–484 17 DNA-Based Detection of Human Pathogenic Fungi 415 Yokoyama K, Biswas SK, Miyaji M, Nishimura K (2000) Identification and phylogenetic relationship of the most common pathogenic Candida species inferred from mitochondrial cytochrome b sequences. J Clin Microbiol 38:4503–4510 Yoshida E, Makimura K, Mirhendi H, Kaneko T, Hiruma M, Kasai T, Uchida K, Yamaguchi H, Tsuboi R (2006) Rapid identification of Trichophyton tonsurans by specific PCR based on DNA sequences of nuclear ribosomal internal transcribed spacer (ITS) 1 region. J Dermatol Sci 42:225–230 Zeng X, Kong F, Halliday C, Chen S, Lau A, Playford G, Sorrell TC (2007) Reverse line blot hybridization assay for identification of medically important fungi from culture and clinical specimens. J Clin Microbiol 45:2872–2880 Zhao J, Kong F, Li R, Wang X, Wan Z, Wang D (2001) Identification of Aspergillus fumigatus and related species by nested PCR targeting ribosomal DNA internal transcribed spacer regions. J Clin Microbiol 39:2261–2266 Zmeili OS, Soubani AO (2007) Pulmonary aspergillosis: a clinical update. QJM 100:317–334 Chapter 18 Applications of Loop-Mediated Isothermal Amplificaton Methods (LAMP) for Identification and Diagnosis of Mycotic Diseases: Paracoccidioidomycosis and Ochroconis gallopava infection Ayako Sano and Eiko Nakagawa Itano Abstract Loop-mediated isothermal amplification (LAMP) methods are now useful for the detection of a specific gene in infectious diseases, genetic diseases, and/ or genetic disorders in the large number of medical fields, and it was recently introduced to fungal investigation. It is characterized by the use of four different primers specifically designed to recognize six distinct regions of the target gene, and the reaction process proceeds at a constant temperature using strand displacement reaction. Quickness and simplicity is the advantage of the method. Amplification and detection of gene can be completed in a single step, by incubating the mixture of samples, primers, DNA polymerase with strand displacement activity and substrates at a constant temperature. The method was applied to two fungal infections; paracoccidioidomycosis (PCM), a deep mycosis caused by Paracoccidioides brasiliensis and Ochroconis gallopava infection. For PCM a combination of F3, B3, FIP, and BIP primers designed from the partial sequence of P. brasiliensis gp43 gene was used. The PCR products amplified by the primer set; F3 and B3 showed species specificity for P. brasiliensis and the detection limit of the PCR was 100 fg of fungal genomic DNA. The specific DNA banding pattern of P. brasiliensis was detected in the clinical and nine-banded armadillo derived isolates, paraffin-embedded tissue sample or sputum from PCM patient. LAMP method was used also for the identification of O. gallopava by using species-specific primer sets based on the D1/D2 domain of the LSU rDNA sequence. The method successfully detected the gene from both fungal DNA derived from brains and spleens of A. Sano Medical Mycology Research Center, Chiba University, 1-8-1, Inohana, Chuo-ku, 260-8673 Chiba, Japan e-mail: aya1@faculty.chiba-u.jp E.N. Itano Department of Pathological Science, CCB, State University of Londrina, P.O. Box 6001, 86051-970 Londrina, Paraná, Brazil e-mail: itanoeiko@hotmail.com Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_18, # Springer-Verlag Berlin Heidelberg 2010 417 418 A. Sano and E.N. Itano experimentally-infected mice with O. gallopava and environmental isolates. In conclusion, LAMP method for PCM and O. gallopava seemed to be useful for identification, diagnosis or retrospective study with advantage in the quickness and simplicity procedure, but require strictly-controlled environments. 18.1 Introduction The gold standard for diagnosis of fungal infections is detection, isolation and identification of the patogenic fungi from clinical specimen; such as skin scrapings, hair and nail, mouth and vaginal swabs, blood, cerebrospinal fluid, urine, sputa and respiratory tract secretion, pus, ocular specimen, and organ biopsy. However, there are many complicated problems in clinical laboratories. First of all, the most important procedure is avoiding laboratory infections caused by highly pathogenic fungal species, such as Coccidioides immitis, C. posadasii, Histoplasma capsulatum, Blastomyces dermatitidis, Paracoccidioides brasiliensis and Penicillium marneffei. In general, highly pathogenic fungal species are not recommended to isolate clinical laboratories in general (Larone 1995). Furthermore, a fungal infection, caused by Ochroconis gallopava, which requires to be differentiated from highly pathogenic bird flu or SARS (Ohori et al. 2006), is also recommended to diagnose without culturing to avoid the viral laboratorial infections. The lower rate of isolation of the causative agents seems to be caused by short incubation periods for the isolation of fungi from clinical material. Visualization of fungal sprouts from clinical materials takes a longer time than those of bacteria. It takes at least 48 h in pathogenic yeasts, 4–7 days in common pathogenic filamentous fungi; dermatophytes, Aspergillus spp., zygomycetes, and others, almost 10–14 days in dematiaceous fungi, and 3 weeks or more in particular fungal species, especially P. brasiliensis, of course, it should not try to isolate in general laboratories. But most laboratories could not keep incubating more than 1 week. Therefore, the isolation rate of filamentous fungal pathogens seems to be very low compared to pathogenic yeasts. Followed by difficulties in isolations of fungal pathogens, there is a serious problem with identification. Expert skills for identification based on morphological and physiological characteristics are strongly required. The procedures are highly time-consuming to prepare photogenic samples and to evaluate special morphologiclal and physiological characteristics (Ohori et al. 2006). Furthermore, clinical isolates sometimes lacked characteristic appearances such as textures and colors of colonies, conidiogenesis, mating and physiological abilities (Uno et al. 2001). Diagnosis for fungal infections is able to be confirmed on the basis of combinations of clinical findings, diagnositic imagings, serological tests, immunological 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 419 tests, cytological and histopathological findings without the fungal isolate (Ishikawa et al. 2008). Detections of 1.3-beta-D-glucan, galactomannan, and D-arabinitol from sera are faster than other methods, and are also useful for diagnosis of some fungal infections (Christensson et al. 1999; Kelaher 2006), however, it is impossible to estimate the species of the causative agent. Isolation and identification, and detection of a species-specific gene by molecular biological methods from clinical materials are able to confirm the causative agent in the species level (Borman et al. 2008). Following the developments of molecular biological techniques in these two decades, molecular biological data for identification of fungal species based on ribosomal DNA (rDNA) sequences became common, such as for P. brasiliensis identification (Motoyama et al. 2000). Although internal transcribed spacer 1 region of rDNA sequence (ITS 1 rDNA) is treated as a barcode gene for identification and taxonomical classification of fungi (Druzhinina 2005), there were some difficulty to confirm the fungal species based on the gene sequences. Therefore, selection of species-specific genes, beside the ITS 1 rDNA may add value to targets. Detections of species-specific genes derived from causative agents in clinical materials by polymerase chain reactions (PCR) have been reported in many mycotic diseases (Reiss et al. 2000; Balajee et al. 2007). But, it still takes at least several hours to obtain the results and sometimes should be requested to confirm the sequences of the amplified genes. Therefore, rapid and accurate diagnostic methods based on molecular biolgical techniques have been waited for. Notomi et al. in 2000 reported a new method, the so called loop-mediated isothermal amplification (LAMP) method to detect specific gene from a DNA virus within a few hours. The method has been applied in the field of microbiology for detection and identification of Mycobacterium sp. (Iwamoto et al. 2003), hepatitis B virus (Nagamine and Watanabe 2001; Nagamine et al. 2002), highly pathogenic bird flu (Imai et al. 2007) and many other pathogens reaching to more than 200 reports up to the end of November 2008. On the other hand, the technique has not been successful in applying fungal infections. Personal communications suggested that the LAMP methods are useful for opportunistic fungal infection in the early times; however, our opinion is that the LAMP methods for Candida spp., Aspergillus spp., and other opportunistic fungal species may not be reliable because of the difficulties to judge the real pathogen or environmental contaminant. We would like to suggest that the application of the LAMP methods to mycotic diseases should be limited to the highly pathogenic fungal species out of endemic areas, and/or to rare species, for example Cryptococcus gattii (Lucas et al. 2009), although a LAMP method for identification of Candida spp. was reported by Inácio et al (2008). The present chapter describes the principle of LAMP method, detections of specific gene from P. brasiliensis categorized as one of the highly pathogenic fungi (Endo et al. 2004), and of O. gallopava required to be differentiated from highly pathogenic bird flu or SARS (Ohori et al. 2006). 420 A. Sano and E.N. Itano FIP 5’ F1c F2 3’ F3 3’ F3 Primer 5’ F3c F2c F1c Targer DNA B1 B2 B3 3’ 5’ 5’ 3’ B1c B2c B3c F3 F2 F1 5’ 3’ B3 B3 primer 3’ B2 5’ B1c BIP Fig. 18.1 Design 4 types of primers based on the following six distinct regions of the target gene: the F3c, F2c and F1c regions at the 30 side and the B1, B2 and B3 regions at the 50 side. FIP: Forward Inner Primer (FIP) consists of the F2 region (at the 30 end) that is complementary to the F2c region, and the same sequence as the F1c region at the 50 end. F3 Primer: Forward Outer Primer consists of the F3 region that is complementary to the F3c region. BIP: Backward Inner Primer (BIP) consists of the B2 region (at the 30 end) that is complementary to the B2c region, and the same sequence as the B1c region at the 50 end. B3 Primer: Backward Outer Primer consists of the B3 region that is complementary to the B3c region (http://loopamp.eiken.co.jp/e/lamp/ primer.html) 18.2 The Principle of LAMP Method 18.2.1 About LAMP Method (http://loopamp.eiken.co.jp/e/ lamp/index.html) LAMP method which stands for LAMP is a simple, rapid, specific and cost-effective nucleic acid amplification method solely developed by Eiken Chemical Co., Ltd. It is characterized by the use of four different primers specifically designed to recognize six distinct regions on the target gene and the reaction process that proceeds at a constant temperature using strand displacement reaction (Fig. 18.1). Amplification and detection of genes can be completed in a single step, by incubating the mixture of samples, primers, DNA polymerase with strand displacement activity and substrates at a constant temperature (about 65 C). It provides high amplification efficiency, with DNA being amplified 109–1010 times in 15–60 min. Because of its high specificity, the presence of amplified product can indicate the presence of the target gene. 18.2.2 Primers LAMP method uses four primer sets; F3, B3, FIP and BIP selected from six distinct regions of the target gene (http://loopamp.eiken.co.jp/e/lamp/primer.html). The most important primer set is F3 and B3. The primers should be selected from specific genes or gene sequences based on species specific PCR and confirmed after 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 421 testing with intra species diversity and a huge numbers of related pathogenic fungal species. Therefore, enormous numbers of trials and errors are latent until the final primers are confirmed. Furthermore, the primers should completely differentiate the fungal genes from host ones. Although, special software to design LAMP primers- PrimerExplore is available in the website (http://primerexplorer.jp/e/), it seemed to be useful as reference hints for base composition, GC contents, secondary structures and Tm value on designing primers based on our experience. 18.2.3 Basic Principle The target gene (DNA template as example) and the reagents are incubated at a constant temperature between 60–65 C. The reaction steps are available at the website (http://loopamp.eiken.co.jp/e/lamp/principle.html). 18.2.4 Cautions LAMP method is highly sensitive. We experienced many faults of contamination of the genes. Once contamination of the target gene occurs, all reactions become positive, even in a negative control using distilled water as a template. Therefore, extremely careful procedures are requested. Reagents, pipets, plastic pipet tips, safety cabinet, and hands. The samples should be handled separately from the reagents. The positive control for the target gene should be done separately. This is one of the reasons why some fungal species common in normal human flora or in environments are not recommended to use the LAMP method. Selection of the target fungal species is also important. The fungal species should be rare in laboratorial environment. 18.3 Applications of LAMP Method for Identifications of P. brasiliensis and or/Diagnosis for Paracoccidioidomycosis (PCM) 18.3.1 Backgrounds for P. brasiliensis P. brasiliensis is considered to belong to the family Onygenaceae (Order Onygenales, Ascomycota), in the same group as Blastomyces dermatitidis, Coccidioides immitis, Histoplasma capsulatum, and Lacazia loboi (Bagagli et al. 2008). The 422 A. Sano and E.N. Itano fungal species is treated as one of the highly pahogenic fungi categorized as biosafety level 3 as the same as C. immitis, C. posadasii, H. capsulatum, B. dermatitidis and Penicillium marneffei (Kamei et al. 2003) On the other hand, the identification and diagnosis of the above fungal infections with nonculture method seems to be very important to avoid laboratory infection (Kamei et al. 2003, Umeyama et al. 2006). P. brasiliensis is the causative agent for paracoccidioidomycosis (PCM) endemic in Latin American countries. This fungus invades the lungs, lymph nodes, skin, mucosa, liver, spleen and various other organs of humans and dogs. In humans, the disease is characterized by two clinical forms: the acute or juvenile form (AF) and more frequently chronic or adult form (CF). The AF is prevalent in children and young people and presents a more severe and rapid clinical evolution with the involvement of multiple organs and adenomegaly, hepatosplenomegaly, digestive disorders, osteo-articular involvement and muco-cutaneous lesions. The CF occurs mainly in adult males and has multiple forms, ranging from benign and localized (unifocal) to severe and disseminated (multifocal) disease that involves skin, mucous membranes, pulmonary and lymph node manifestations (Restrepo 1985; Franco 1987; Kwon-Chung and Bennett 1992; Brummer et al. 1993; Ono et al. 2003; Ricci et al. 2004). The probable natural habitat of P. brasiliensis is soil as saprophytic form. In fact, isolations from soil or soil related products, from the feces of both frugivorous bats (Artibeus lituratus) and a penguin (Pygoscelis adeliae) were reported. Interestingly, the natural reservoir of P. brasiliensis seems to be the nine-banded armadillo (Dasypus novemcinctus) because of repeated isolation of the fungal species from various endemic areas of paracoccidioidomycosis with high incidences showing, as the same genetic profiles as clinical isolates. Furthermore, detection of P. brasiliensis gene from the internal organs of wild animals that died in traffic accidents; guinea pig (Cavia aperea), porcupine (Sphiggurus spinosus), grison (Gallictis vittata) and raccoon (Procyon cancrivoros) suggested that the mycosis invades not only humans but also many mammal species, and is one of zoonotic mycosis (Bagagli et al. 2008). The characteristics of P. brasiliensis is temperature-dependent dimorphism; a mycelial form at ambient temperature, and multiple budding yeast form in host tissue or at temperatures above 35–37 C in certain culture media (Restrepo 1985; Franco 1987; Franco et al. 1989; Kwon-Chung and Bennett 1992; Brummer et al. 1993). (Fig. 18.2). The important characteristics of P. brasiliensis are a multiple-nuclei microorganism (Imai et al. 2000), and a haploid microorganism except for gp43, encoding the major antigen of P. brasiliensis; 43 kDa glycoprotein appeared as only one allele (Almeida et al. 2007). The mature or during maturation of yeast cells that have multiple nuclei, suggested that any P. brasiliensis gene may easily be amplified because of multiple copies at least. It suggested that selection of the target gene is free from sticking on ribosomal RNA genes having tandem repeats (Kobayashi 2006) from clinical materials. 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 423 Fig. 18.2 (a) Upper; colony of P. brasiliensis cultured on Sabouraud dextrose agar plate, lower; on potato dextrose agar plate at 25 C for 8 weeks, (b) cerebriform yeast-like colony on 1% dextrose added brain heart infusion agar slant cultured at 35 C for 7 days (left; ninebanded armadillo derived isolate, right; clinical isolate), (c) whip wheel like appearance in a infected lymphonode tissue (Dr. Nakajima Y, Matsushita Memorial Hospital, Osaka, Japan), (d) aleurioconidia cultured on potato dextrose agar at 25 C for 8 weeks, (e) clamydospores cultured on potato dextrose agar at 25 C for 8 weeks, (f) mycelial to yeast form conversion process cultured on potato dextrose agar at 25 C for 2 weeks and cultured at 35 C for 3 days, (g) multiple budding yeast cells consisted of big mother cells, daughter and grand daughter cells Genetic data of P. brasiliensis have progressed in the twenty first century. More than 5,000 sequences of P. brasiliensis are released into the GenBank database (http://www.ncbi.nlm.nih.gov/sites/entrez). Based on multiple gene analysis, P. brasiliensis was separated into three different phylogenetic species; S1 (species 1 from Brazil, Argentina, Paraguay, Peru and Venezuela), PS2 (phylogenetic species 2 from Brazil and Venezuela) and PS3 (phylogenetic species 3 from Colombia) (Matute et al. 2006a, b, 2007). Whole genome sequences on three strains of P. brasiliensis (Pb01, Pb03 and Pb18) were released in the BROAD Institute (http://www.broad.mit.edu/annotation/ genome/paracoccidioides_brasiliensis/MultiHome.html). According to Carrero et al. 2008, isolate Pb01 might be a new Paracoccidioides species because of its diversity of gene profiles compared to other P. brasiliensis isolates, and was named as P. lutzii (Teixeira et al. 2009). Among various genes, the gp43 is the most important gene because of its diagnostic value (Puccia et al. 2008). We have also been trying to detect gp43 from paraffin embedded tissue samples and blood (Sano et al. 2001; Itano et al. 2002). 424 A. Sano and E.N. Itano The gene encodes the major fungal antigen; 43 kDa glycoprotein, which is a dominant P. brasiliensis antigen, and has been used for serological test in endemic areas (Miura et al. 2001; Camargo 2008). Approximately 300 sequences of gp43 were released in the GenBank database at the end of August 2009. The gene homologies among the majority of P. brasiliensis isolates is more than 96% in, except for Pb01 and its related isolates identity (Teixeira et al. 2009, Takayama et al. 2009). According to Takayama et al., the LAMP band pattern of P. lutzii was different from those of P. brasiliensis. 18.3.2 LAMP Method for Identifications of P. brasiliensis 18.3.2.1 P. brasiliensis Isolates and Reference Species Twenty-two clinical and seven nine-banded armadillo (Dasypus novemcinctus) derived P. brasiliensis isolates were tested. As an advanced notice, our method might limit to detect S1 phylogenetic type of P. brasiliensis since we have not tested the isolates belonging to PS2 and PS3 proposed by Matute et al. 2006a and b. Furthermore, there is an uncertainty to detect gp43 in atypical isolate P. brasiliensis strain Pb01 and its related isolates (Carrero et al. 2008; Takayama et al. 2009), and has just been named as the new species P. lutzii (Teixeira et al. 2009). Isolates of Coccidioides immitis sensu lato (IFM 50993, identified as C. posadasii based on multiple gene analysis by Sano et al. 2006), Histoplasma capsulatum (IFM 41329), Blastomyces dermatitidis (IFM 41316), Sporothrix schenckii (IFM 47068), Penicillium marneffei (IFM41708), Candida albicans (IFM 5740), and Cryptococcus neoformans (IFM 5830) were used as negative controls (Table 18.1). 18.3.2.2 Extraction of DNA Isolates of P. brasiliensis were evaluated. Yeast-form cells harvested on 1.0% glucose added DifcoTM brain heart infusion agar (Becton Dickinson Microbiology Systems, Sparks, MD, USA) slants at 35 C for 7 days were used. Approximately 5  108 yeast-form cells were suspended in distilled water (DW) and washed three times with DW, and homogenized in a 1.5 mL volume plastic homogenizer. DNA was extracted with the Gen Toru Kun for the yeast (Dr. GenTLETM for yeast) kit (TAKARA BIO INC., Ohtsu, Shiga, Japan). Isolates of C. posadasii H. capsulatum, B. dermatitidis, S. schenckii, P. marneffei, Ca.albicans, and Cr. neoformans were cultured on potato dextrose agar (Becton Dickinson Microbiology Systems) at 25 C for 7–60 days. The fungal cells of C. posadasii were fixed with 70% ethanol overnight, and the DNA was extracted by the kit (Dr. GenTLETM for yeast, TAKARA BIO INC.). The final concentrations of DNA were adjusted from 10 to 20 ng/mL. Country (City) Source (Remarks) Phylogenetic speciesa Accession no. (gp43) Pb-9 Pb-18 Bt-2 Bt-3 Bt-4 Bt-7 Bt-9 B1183 PbLev B339 Brazil Brazil Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Brazil Brazil Brazil Human patient Human patient Human patient Human patient Human patient Human patient Human patient Human patient Human patient Human patient S1 S1 S1 S1 S1 S1 S1 ND S1 S1 AB047690 AB047691 AB304676 AB304677 AB047693 AB304678 AB047694 ND AB304680 AB304681 Recife Pb-HM-AOK Hachisuga WAG Tateishi Bt-1 Pb-267 Pb-265 Recife-Pb-HC P-25 P-30 UMK Tatu PRT1 PRT2 D3LY1 D4S1 Brazil (Recife) Japan (Tokyo)b Japan (Fukuoka)b Japan (Osaka)c Japan (Ibaragi)b Brazil (Botucatu, São Paulo) Brazil Brazil Brazil (Recife) Costa Rica (San Jose) Costa Rica (San Jose) Japan (Chiba)b Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Human patient Human patient Human patient Human patient Human patient Human patient Mutant of Pb-9 Mutant of Pb-9 Human patient Human patient Human patient Human patient Armadillo Armadillo Armadillo Armadillo Armadillo S1 S1 S1 S1 S1 S1 S1 S1 S1 ND ND S1 S1 S1 S1 S1 S1 AB304682 AB047695 AB304682 AB047696 AB304684 AB304685 AB047692 AB304686 AB047699 AB047698 AB304688 AB047697 AB047700 AB047701 AB047702 AB047813 AB047704 (continued) 425 Strain Applications of Loop-Mediated Isothermal Amplificaton Methods (=CBS 372.73, =ATCC 32069) IFM 41631 IFM 41632 IFM 41633 IFM 46215 IFM 46240 IFM 46464 IFM 46465 IFM 46466 IFM 46467 IFM 46468 IFM 46470 IFM 46930 IFM 46463 IFM 47183 IFM 47185 IFM 47195 IFM 47217 18 Table 18.1 Isolates IFM Number Paracoccidioides brasiliensis IFM 41620 IFM 41621 IFM 41622 IFM 41623 IFM 41624 IFM 41625 IFM 41626 IFM 41628 IFM 41629 IFM 41630 Country (City) Brazil (Botucatu, São Paulo) Brazil (Botucatu, São Paulo) Source (Remarks) Armadillo Armadillo Phylogenetic speciesa S1 S1 Accession no. (gp43) AB047703 AB047705 Coccidioides immitis sensu lato (C. posadasii) IFM 50993 USA Human patient – – Histoplasma capsulatum IFM 41329 USA Human patient – – USA Human patient – – Sporothrix schenkii IFM 47068 Japan Human patient – – Penicillium marneffei IFM 41708 China Bamboo rat – – Candida albicans IFM 5740 Japan Human patient – – Blastomyces dermatitidis IFM 41316 (=ATCC 26199) Strain D4S9 D4LIV1 A. Sano and E.N. Itano Cryptococcus nenformans sensu lato IFM 5830 Japan Human patient – – IFM Institute of Food Microbiology, Chiba University, the former name of the Medical Mycology Researc Center, and deposited as the official abbreviation of the world culture collection of pathogenic fungi and actinomycetes a Phylogenetic species was estimated from gp43 sequence b The patient was infected in Brazil c The patient was infected in Paraguay ND Not determined 426 Table 18.1 (continued) IFM Number IFM 47228 IFM 47247 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 427 DNA extracted from a paraffin-embedded tissue sample of PCM and an ethanolfixed sputum sample was extracted with a DEXPAT kit (TAKARA BIO INC.) and was also used in the LAMP assay. 18.3.2.3 Detection of gp43 by PCR A total volume of 25 mL was used for all PCR reactions. Fifty nanograms per milliliter of DNA extracts were added to 2.5 mL of Ex TaqTM buffer in the kit (Ex TaqTM, TAKARA BIO INC.) containing 4.5 mM MgSO4, 2 mL (2.5 mM each) dNTP mixture in the kit (ExTaqTM, TAKARA BIO INC.), 2 mL each 10 pM primer set of F3 50 -TCA CGT CGC ATC TCA CAT TG-30 encoding from 391st to 410th and B3 50 -AAG CGC CTT GTC CAA ATA GTC GA-30 designed from the complementary sequence from 718th to 740th correspondent to gp43 sequence at GenBank U26160 and 0.0625 mL (5 units/mL) TaKaRa Ex TaqTM polymerase in the kit (Ex TaqTM, TAKARA BIO INC.). Reaction mixtures were subjected to denaturation at 94 C for 1 min, followed by 30 cycles of amplification at 94 C for 1 min, 50 C for 1 min, and 72 C for 2 min and a final extension at 72 C for 10 min, in a PCR Thermal Cycler MP (TAKARA BIO INC.). PCR products were separated by electrophoresis on 1.0% agarose gels in TAE buffer (40 mM Tris-base, 20 mM acetic acid, 1 mM EDTA), stained with ethidium bromide, and visualized by UV transillumination. DNA strands obtained from the PCR were processed for direct sequencing with ABI Prism 3,100 (Applied Biosystems, Foster City, CA., USA) to confirm the sequence of gp43 (Sano et al. 1998–1999). 18.3.2.4 LAMP Method for gp43 Briefly, the LAMP method used in the present study detects the gp43 gene with a combination of F3, B3, FIP, and BIP primers designed from the partial sequence of gp43 (GenBank accession number U26160) by a registration system primer designing website (FUJITSU Ltd., Tokyo, Japan: “LAMP PIMER EXPLORER” website in “Netlaboratory” homepage http://venus.netlaboratory.com/partner/lamp/index. html). These primers recognize a partial sequence of gp43. The primer sequences were as follows: F3, used in the species specific forward primer; B3, used in the species-specific reverse one; FIP, 50 -TGG CTC CAG CAA TAG CCA CCC GTC AAG CAG GAT CAG CAA T-30 designed from the forward sequence of 425th to 445th and the complementary sequence of 464th to 485th; and BIP: 50 -CAT GTC AGG ATC CCG ATC GGG CCT TGT ACA TAT GGC TCT CCC T-30 designed by the forward sequence from 648th to 668th and the complementary sequence from 691st to 712th. The annealing sites of the primers are shown in Fig. 18.3. One micro liter of 10 ng/mL DNA template and 40 pmol each of the FIP and BIP primers and 5 pmol each of the F3 and B3 primers were mixed with 12.5 mL of 2 reaction mix in the kit (Loop AMP, Eiken Chemical Co., Ltd., Tokyo, Japan) in a 428 A. Sano and E.N. Itano 1981 1 391 F3 Primer (391–410)f FIP Primer (425–445)f +(485–464)c BIP Primer (648–668)f +(712–691)c 740 B3 Primer (740–718)c Fig. 18.3 Primer map for the LAMP method of detecting gp43 from P. brasiliensis final volume of 23.0 mL. DNA mixtures were incubated at 63 C for 60 min. The reaction was stopped by heating the mixture at 80 C for 2 min to inactivate the enzyme of LAMP amplification. Detection limits of the LAMP method were evaluated with serial dilutions of DNA from isolate IFM 46930. As the positive control attached with the kit and a negative control consisted of DW and other fungal DNAs, C. immitis, H. capsulatum, B. dermatitidis, S. schenckii, P. marneffei, C. albicans, and Cr. neoformans were used. In addition, DNAs extracted from a paraffin-embedded tissue sample, and an ethanol-fixed sputum were reacted at 63 C for 60 and 120 min. In addition, time-dependent increases in levels of DNA products by LAMP were monitored by real-time-PCR (Rotor-Gene, RG2000, NIPPN/Techno Cluster, Inc., Tokyo, Japan) for as long as 70 min at 63 C with P. brasiliensis isolates IFM 41630 and IFM 46215. 18.4 Results The PCR products amplified with the primer set; F3 and B3 showed species specificity for P. brasiliensis. The detection limit of the PCR was 100 fg of fungal genomic DNA (data not shown). Other related species, such as C. posadasiis (not shown), H. capsulatum, and B. dermatitidis or important pathogenic fungi; S. schenckii, P. marneffei, Ca. albicans, and Cr. neoformans were negative (Fig. 18.4). All partial sequences of gp43 consisted of 339 bps and were correspondent to their accession numbers, except for isolate IFM 41628 (not done). The specific DNA banding pattern of P. brasiliensis was detected in the clinical and nine-banded armadillo derived isolates by LAMP. No DNA band was observed in negative control isolates of C. posadasii, H. capsulatum, B. dermatitidis, S. schenckii, P. marneffei, C. albicans, and Cr. neoformans (Fig. 18.5). The detection limit of LAMP for gp43 was also 100 fg of fungal genomic DNA. The incubation procedure at 63 C for 60 min was not sufficient for detection of gp43 from DNA extracted from paraffin-embedded tissue sample or sputum infected with PCM (data not shown). The DNA from a paraffin-embedded tissue and sputum from different patients yielded the same ladder band yielded by fungal DNAs via LAMP at 63 C for 120 min (Fig. 18.6a, b). 18 Applications of Loop-Mediated Isothermal Amplificaton Methods bps 429 bps 1000 1000 500 500 M 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 M Fig. 18.4 Amplification of the gp43 gene by PCR with primers F3 and B3. All DNA derived from P. brasiliensis isolates (line 1–11) were uniformly positive. 12: Ca. albicans, 13: H. capsulatum, 14: B. dermatitidis, 15: P. marneffei, 16: S. schenckii, 17: Cr. neoformans, and 18: C. immitis (C. posadasii) were negative bps bps 1000 1000 500 500 M 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 M Fig. 18.5 Amplification of the gp43 gene by the LAMP methods. All DNA derived from P. brasiliensis isolates (line 1–11) were uniformly positive. 12: Ca. albicans, 13: H. capsulatum, 14: B. dermatitidis, 15: P. marneffei, 16: S. schenckii, 17: Cr. neoformans, and 18: C. immitis 18: C. immitis (C. posadasii) were negative The LAMP reaction reached a plateau after incubation at 63 C for 45 min, so far, as monitored by real time- PCR (Fig. 18.7). The positive control provided with the kit reached a plateau at 15 min, and the negative one did not show increase of fluorescence level. DNAs from other fungal species did not increase the fluorescence level (data not shown). The LAMP reaction of DNA from isolate IFM 46215 reached a plateau at 63 C for 45 min and those of IFM 41622 was 50 min. 430 A. Sano and E.N. Itano Fig. 18.6 (a) Amplification of the gp43 from paraffin embedded tissue sample by the LAMP methods. M: Marker, 2: DNA from the paraffin embedded tissue sample. 3 and 4: Fungal DNA of P. brasiliensis. (b) Those from sputa. M: marker, 2: DNA from the sputum, 3 and 4: Fungal DNA of P. brasiliensis bps a bps 1000 b 1000 M 1 2 3 M M 1 2 3 35 30 Fluorescence 25 20 15 Positive control P. brasiliensis (IFM 41622) P. brasiliensis (IFM 46215) Negative control (DW) 10 5 0 0 10 20 30 Minutes 40 50 60 Cycle Fig. 18.7 LAMP reaction monitored by real-time-PCR. The negative control with other fungal DNAs, C. immitis (C. posadasii), H. capsulatum, B. dermatitidis, S. schenckii, P. marneffei, Ca. albicans, and Cr. neoformans were as the same as DW 18.5 Comments and Opinions The LAMP method provides for more rapid detection of gp43 than nested PCR. LAMP required only 3 h from DNA extraction to identification, whereas nested PCR required 12 h when we tested. 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 431 LAMP methods are also advantageous because it can be applied to clinical material, such as paraffin-embedded tissue and sputum samples for retrospective study (Endo et al. 2004; Tatibana et al. 2009). Even in clinical samples, the time required for diagnosis was less than 4 h. The LAMP method is not only convenient for identification of P. brasiliensis, but also for diagnosis of PCM, especially for identification of P. brasiliensis and diagnosis of PCM outside of the endemic areas, such as European countries and Japan. Patients in endemic areas are sometimes misdiagnosed as having a malignant tumor because of a shadow on the chest X-ray and granulomatous inflammation of infected tissue. Therefore, most PCM of patients in Japan are being diagnosed on the basis of histopathological findings (Endo et al. 2004). The LAMP method could be applied for PCM diagnosis in such cases without isolation of the fungus. Application of real-time-PCR to the LAMP method should shorten the time for obtaining the results within a couple of hours, because electrophoresis is not required. While analysis of LAMP amplification products by agarose gel electrophoresis takes approximately 3 h, LAMP in connection with real-time-PCR takes only 2 h. According to the manufacturer’s protocol, LAMP products can be detected by optical density under UV light. However, we do not recommend this method. Some of pseudo reactions showing smear-like amplification products also became positive. Furthermore, we do not have any experience to react as a smear-like amplification in the real-time PCR method. Uncertainty of the reaction also could not be removed. Therefore, LAMP products should be visualized by agarose gel electrophoresis. In addition, the reaction does not require a special thermo cycler system. A styrofoam box with warm water like that of a hot coffee temperature is one sign of a good apparatus. It suggested that the method is useful in field hospitals. This method will be important for detecting specific genes in highly pathogenic or rare emerging fungal infections which require care and time-consuming culturing procedures. However, because of extremely higher sensitivities to detecting genes by the LAMP methods, it should be meaningless to apply the LAMP methods to Candida species that exist as common fungal flora in oral or body surface, to Aspergillus species and/or to other causative agents for the emerging fungal infections habitat popular in soil or environments. It should be impossible to judge the results whether it is environmental contaminations or real infectious propague. In addition, we would like to avoid to give a comment on the report by Inácio et al (2008). 18.6 LAMP Method for Identifications of O. gallopava We also applied the LAMP method to detection of the species-specific gene of O. gallopava; a species of dematiaceous fungi recognized as a causative agent of zoonotic and emerging fungal infections. The fungal specie shows excellent growth at 42 C (Fig. 18.8a), and is able to grow up to 45 C or more. 432 A. Sano and E.N. Itano Fig. 18.8 (a) Colonies of O. gallopava cultured on potato dextrose agar plate at 25, 37 and 42 C for 8 days, (b and c) clavate conidia under microscopy, x400 It affects the central nervous system and respiratory tracts of humans, birds and cats and is required to be differentiated from SARS and highly pathogenic bird flu. Clavate conidia (Fig. 18.8b and c) are virulent to experimentally infected mice (Ohori et al. 2006; Yarita et al. 2007). We designed O. gallopava species-specific primer sets to aid in its identification by the LAMP method based on the D1/D2 domain of the LSU rDNA sequence. The primer set for O. gallopava was designed based on the sequence of D1/D2 LSU rDNA of O. gallopava (accession number AB125281 in GenBank) with a comparison of 21 species of dematiaceous fungi obtained from the present study and from 108 sequences in GenBank database. The primer sequences were as follows: OgF3: 50 -AGG GAG TCT CGG GTT AAG GG-30 encoding from the 391st to the 410th, and OgB3: 50 -CAT TCC CTT CGT CTT TGT CC-30 corresponding to the complementary sequence from the 718th to the 740th of AB125281 and were species-specific for O. gallopava (Fig. 18.9). FIP; 50 -ACT CGA CTC GTC GAA GGG GCA GAG GGT GAG AGT CCC GT-30 designed by the forward sequence of 425th to 445th and the complementary sequence of 464th to 485th, and BIP; 50 -ACT GGC CAG AGA CCG ATA GCG TGA CTC TCT TTT 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 433 bps bps 1000 1000 500 500 M 1 2 3 4 5 6 7 8 9 10 11 12 13 M Fig. 18.9 Species specific PCR for O. gallopava. M: Marker, 1–10: O. gallopava, 11: O. gamsii, 12: and 13: O. tsawytschae. The related species such as O. constricta, O. humicola, Alternaria alternata, Arthrobotrys javanica, Bipolaris sp., Bipolaris specifera, Cladophialophora bantiana, C. carrionii, Curvularia geniculata, Cu. lunata var. lunata, Cu. senegalensis, Exophiala alcalophiala, E. dermatitidis, E. jeanselmei, E. moniliae, E. spinifera, Fonsecaea pedrosoi, Phialophora verrucosa, Rhinocladiella atrovirens, Scolecobasidium terreum were negative bps bps 1000 1000 500 500 M 1 2 3 4 5 6 7 8 9 10 11 12 13 M Fig. 18.10 Species specific loop mediated isothermal amplification method (LAMP) for O. gallopava. M: Marker, 1–10: O. gallopava, 11: O. gamsii, 12: and 13: O. tsawytschae. The related species such as O. constricta, O. humicola, Alternaria alternata, Arthrobotrys javanica, Bipolaris sp., Bipolaris specifera, Cladophialophora bantiana, C. carrionii, Curvularia geniculata, Cu. lunata var. lunata, Cu. senegalensis, Exophiala alcalophiala, E. dermatitidis, E. jeanselmei, E. moniliae, E. spinifera, Fonsecaea pedrosoi, Phialophora verrucosa, Rhinocladiella atrovirens, Scolecobasidium terreum were negative CAA AGT GC-30 designed by the forward sequence from 648th to 668th and the complementary sequence from 691 st to 712 nd of AB125281. The LAMP method successfully detected the gene from both the fungal DNA derived from experimentally infected brains and spleens of mice and environmental 434 A. Sano and E.N. Itano bps bps 1000 1000 500 500 M 1 2 3 4 5 6 7 8 9 10 11 12 13 M Fig. 18.11 Detection of O. gallopava gene from the experimentally infected brains and spleens of mice by LAMP method. M: Marker, 1–5: brain tissue of mice infected with O. gallopava. 6: blank, 7–11: spleen tissue of mice infected with O. gallopava. 12 and 13: negative control DNA from demateaceous fungi bps Fig. 18.12 DNA pattern by loop madiated isothermal amplification method (LAMP) specific for O. gallopava using 20 pg of fungal DNA. M: Marker, 1 and 6: a clinical isolate, 2–5: hot spring isolates, 7: a negative control using distilled water for a template bps 1000 1000 500 500 M 1 2 3 4 5 6 7 M isolates (Fig. 18.10–18.12),which will help to differentiate O. gallopava infection from other important avian zoonoses (Ohori et al. 2006; Yarita et al. 2007). 18.7 Conclusion and Future Line of Research In conclusion, LAMP method for PCM and O. gallopava seemed to be useful for fungal identification, diagnosis or retrospective study with advantage in the quickness and simplicity procedure, but require strictly controlled environments. It could be applicable for clinical identification of fungi and diagnosis of fungal 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 435 diseases caused by level 3 biohazards, such as coccidioidomycosis, histoplasmosis, blastomycosis, and infection of Penicillium marneffei, which generally require care and time consuming culturing procedures, and causative agent for emerging fungal infections. Acknowledgments We thank Drs. Kazuko Nishimura, Makoto Miyaji, Shigeo Endo, Akira Ohori, Tsuyoshi Igarashi, Koji Yokoyama, Masashi Yamaguchi, Yoko Takahashi, Katsuhiko Kamei, Marcello Franco, Giannina Ricci, Berenice Tatibana, Ms. Kyoko Yarita and Mr. Takashi Komori for their cooperation. References Almeida AJ, Matute DR, Carmona JA, Martins M, Torres I, McEwen JG, Restrepo A, Leão C, Ludovico P, Rodrigues F (2007) Genome size and ploidy of Paracoccidioides brasiliensis reveals a haploid DNA content: flow cytometry and GP43 sequence analysis. Fungal Genet Biol 44:25–31 Bagagli E, Theodoro RC, Bosco SM, McEwen JG (2008) Paracoccidioides brasiliensis: phylogenetic and ecological aspects. Mycopathol 165:197–207 Balajee SA, Sigler L, Brandt ME (2007) DNA and the classical way: identification of medically important molds in the 21st century. Med Mycol 45:475–490 Borman AM, Linton CJ, Miles SJ, Johnson EM (2008) Molecular identification of pathogenic fungi. J Antimicrob Chemother 61(Suppl 1):i7–i12 Brummer E, Castaneda E, Restrepo A (1993) Paracoccidioidomycosis: an update. Clin Microbiol Rev 6:89–117 Camargo ZP (2008) Serology of paracoccidioidomycosis. A centennial: discovery of Paracoccidioides brasiliensis. Mycopathol 165:289–302 Carrero LL, Niño-Vega G, Teixeira MM, Carvalho MJ, Soares CM, Pereira M, Jesuino RS, McEwen JG, Mendoza L, Taylor JW, Felipe MS, San-Blas G (2008) New Paracoccidioides brasiliensis isolate reveals unexpected genomic variability in this human pathogen. Fungal Genet Biol 45:605–612 Christensson B, Sigmundsdottir G, Larsson L (1999) D-arabinitol–a marker for invasive candidiasis. Med Mycol 37:391–396 Druzhinina IS (2005) An oligonucleotide barcode for species identification in Trichoderma and Hypocrea. Fungal Genet Biol 42:813–828 Endo S, Komori T, Ricci G, Sano A, Yokoyama K, Ohori A, Kamei K, Franco M, Miyaji M, Nishimura K (2004) Detection of gp43 of Paracoccidioides brasiliensis by the loop-mediated isothermal amplification (LAMP) method. FEMS Microbiol Lett 234:93–97 Franco M (1987) Host-parasite relationships in paracoccidioidomycosis. J Med Vet Mycol 25:5–18 Franco M, Sano A, Kera K, Nishimura K, Takeo K, Miyaji M (1989) Chlamydospore formation by Paracoccidioides brasiliensis mycelial form. Rev Inst Med Trop Sao Paulo 31:151–157 Imai T, Sano A, Mikami Y, Watanabe K, Aoki FH, Branchini ML, Negroni R, Nishimura K, Miyaji M (2000) A new PCR primer for the identification of Paracoccidioides brasiliensis based on rRNA sequences coding the internal transcribed spacers (ITS) and 5.8S regions. Med Mycol 38:323–326 Imai M, Ninomiya A, Minekawa H, Notomi T, Ishizaki T, Van Tu P, Tien NT, Tashiro M, Odagiri T (2007) Rapid diagnosis of H5N1 avian influenza virus infection by newly developed influenza H5 hemagglutinin gene-specific loop-mediated isothermal amplification method. J Virol Methods 141:173–180 436 A. Sano and E.N. Itano Inácio J, Flores O, Spencer-Martins I (2008) Efficient identification of clinically relevant Candida yeast species by use of an assay combining panfungal loop-mediated isothermal DNA amplification with hybridization to species-specific oligonucleotide probes. J Clin Microbiol 46:713–720 Ishikawa H, Oda S, Murata A, Shimazaki S, Hirasawa H, Aikawa N (2008) Influence of the diagnosis and treatment guidelines for mycosis profunda (deep mycosis) in the field of emergency and critical care medicine–with reference to patient background. Jpn J Antibiot 61:18–28 (In Japanese) Itano E, Uno J, Sano A, Yarita K, Kamei K, Miyaji M, Nishimura K, Mikami Y (2002) 506 Detection of the gp43 gene and (1–3)-beta-D-glucan of Paracoccidioides brasiliensis in the 507 blood of experimentally infected mice. Nippon Ishinkin Gakkai Zasshi 43:29–35 Iwamoto T, Sonobe T, Hayashi K (2003) Loop-mediated isothermal amplification for direct detection of Mycobacterium tuberculosis complex, M. avium, and M. intracellulare in sputum samples. J Clin Microbiol 41:2616–2622 Kamei K, Sano A, Kikuchi K, Makimura K, Niimi M, Suzuki K, Uehara Y, Okabe N, Nishimura K, Miyaji M (2003) The trend of imported mycoses in Japan. J Infect Chemother 9:16–20 Kelaher A (2006) Two non-invasive diagnostic tools for invasive aspergilosis: (1–3)-beta-Dglucan and the galactomannan assay. Clin Lab Sci 19:222–224 Kobayashi T (2006) Strategies to maintain the stability of the ribosomal RNA gene repeats– collaboration of recombination, cohesion, and condensation. Genes Genet Syst 81:155–161 Kwon-Chung KJ, Bennett JE (1992) In: Kwon-Chung KJ, Bennett JE (eds) Medical mycology. Pa: Lea & Febiger, Philadelphia, pp 594–619 Larone DH (1995) Safety precautions. In: Larone DH (ed) Medically important fungi, 3rd edn. ASM press, Washington D.C., USA, pp 5–6 Lucas S, da Luz Martins M, Flores O, Meyer W, Spencer-Martins I, Inácio J (2009) Differentiation of Cryptococcus neoformans varieties and Cryptococcus gattii using CAP59-based loopmediated isothermal DNA amplification. Clin Microbiol Infect Aug 20. [Epub ahead of print] Matute DR, McEwen JG, Puccia R, Montes BA, San-Blas G, Bagagli E, Rauscher JT, Restrepo A, Morais F, Niño-Vega G, Taylor JW (2006a) Cryptic speciation and recombination in the fungus Paracoccidioides brasiliensis as revealed by gene genealogies. Mol Biol Evol 23:65–73 Matute DR, Sepulveda VE, Quesada LM, Goldman GH, Taylor JW, Restrepo A, McEwen JG (2006b) Microsatellite analysis of three phylogenetic species of Paracoccidioides brasiliensis. J Clin Microbiol 44:2153–2157 Matute DR, Torres IP, Salgado-Salazar C, Restrepo A, McEwen JG (2007) Background selection at the chitin synthase II (chs2) locus in Paracoccidioides brasiliensis species complex. Fungal Genet Biol 44:357–367 Miura CS, Estevão D, Lopes JD, Itano EN (2001) Levels of specific antigen (gp43), specific antibodies and antigen-antibody complexes in saliva and serum of paracoccidioidomycosis patients. Med Mycol 39:423–428 Motoyama AB, Venancio EJ, Brandão GO, Petrofeza-Silva S, Pereira IS, Soares CM, Felipe MS (2000) Molecular identification of Paracoccidioides brasiliensis by PCR amplification of ribosomal DNA. J Clin Microbiol 38:3106–3109 Nagamine K, Hase T, Notomi T (2002) Accelerated reaction by loop-mediated isothermal amplification using loop primers. Mol Cell Probes 16:223–229 Notomi T, Okayama H, Masubuchi H, Yonekawa T, Watanabe K, Amino N, Hase T (2000) Loopmediated isothermal amplification of DNA. Nucleic Acids Res 28:E63–e63 Ohori A, Endo S, Sano A, Yokoyama K, Yarita K, Yamaguchi M, Kamei K, Miyaji M, Nishimura K (2006) Rapid identification of Ochroconis gallopava by a loop-mediated isothermal amplification (LAMP) method. Vet Microbiol 114:359–365 Ono MA, Kishima MO, Itano EN, Bracarense AP, Camargo ZP (2003) Experimental paracoccidioidomycosis in dogs. Med Mycol 41:265–268 Puccia R, McEwen JG, Cisalpino PS (2008) Diversity in Paracoccidioides brasiliensis. The Pb GP43 gene as a genetic marker. A centennial: discovery of Paracoccidioides brasiliensis. Mycopathol 165:275–287 18 Applications of Loop-Mediated Isothermal Amplificaton Methods 437 Reiss E, Obayashi T, Orle K, Yoshida M, Zancopé-Oliveira RM (2000) Non-culture based diagnostic tests for mycotic infections. Med Mycol 38(Suppl 1):147–159 Restrepo AM (1985) The ecology of Paracoccidioides brasiliensis: a puzzle still unsolved. J Med Vet Mycol 23:323–334 Ricci G, Mota FT, Wakamatsu A, Serafim RC, Borra RC, Franco M (2004) Canine paracoccidioidomycosis. Med Mycol 42:379–383 Sano A, Defaveri J, Tanaka R, Yokoyama K, Kurita N, Franco M, Coelho KI, Bagagli E, Montenegro MR, Miyaji M, Nishimura K (1998–1999) Pathogenicities and GP43kDa gene of three Paracoccidioides brasiliensis isolates originated from a nine-banded armadillo (Dasypus novemcinctus). Mycopathology 144:61–65 Sano A, Yokoyama K, Tamura M, Mikami Y, Takahashi I, Fukushima K, Miyaji M, Nishimura K (2001) Detection of gp43 and ITS1–5.8S-ITS2 ribosomal RNA genes of Paracoccidioides brasiliensis in paraffin-embedded tissue. Nippon Ishinkin Gakkai Zasshi 42:23–27 Sano A, Miyaji M, Kamei K, Mikami Y, Nishimura K (2006) Reexamination of Coccidioides spp. reserved in the Research Center for Pathogenic Fungi and Microbial Toxicoses, Chiba University, based on a multiple gene analysis. Nippon Ishinkin Gakkai Zasshi 47:113–117 Takayama A, Itano EN, Sano A, Ono MA, Kamei K (2009) An atypical Paracoccidioides brasiliensis clinical isolate based on multiple gene analysis. Med Mycol 2009 Feb 19:1–9 [Epub ahead of print] Tatibana BT, Sano A, Uno J, Kamei K, Igarashi T, Mikami Y, Miyaji M, Nishimura K, Itano EN (2009) Detection of Paracoccidioides brasiliensis gp43 gene in sputa by loop-mediated isothermal amplification method. J Clin Lab Anal 23:139–143 Teixeira MM, Theodoro RC, de Carvalho MJ, Fernandes L, Paes HC, Hahn RC, Mendoza L, Bagagli E, San-Blas G, Felipe MS (2009) Phylogenetic analysis reveals a high level of speciation in the Paracoccidioides genus. Mol Phylogenet Evol 52:273–283 Umeyama T, Sano A, Kamei K, Niimi M, Nishimura K, Uehara Y (2006) Novel approach to designing primers for identification and distinction of the human pathogenic fungi Coccidioides immitis and Coccidioides posadasii by PCR amplification. J Clin Microbiol 44:1859–1862 Uno J, Tanaka R, Branchini ML, Aoki FH, Yarita K, Sano A, Fukushima K, Mikami Y, Nishimura K, Miyaji M (2001) Atypical Cryptococcus neoformans isolate from an HIV-infected patient in Brazil. Nippon Ishinkin Gakkai Zasshi 42:127–132 Nagamine K, Watanabe K (2001) Loop-mediated isothermal amplification reaction using a nondenatured template. Clin Chem 47:1742–1743 Yarita K, Sano A, Murata Y, Takayama A, Takahashi Y, Takahashi H, Yaguchi T, Ohori A, Kamei K, Miyaji M, Nishimura K (2007) Pathogenicity of Ochroconis gallopava isolated from hot springs in Japan and a review of published reports. Mycopathology 164:135–147 Chapter 19 Identification of the Genus Absidia (Mucorales, Zygomycetes): A Comprehensive Taxonomic Revision Kerstin Hoffmann Abstract This brief review intends to survey and evaluate the present knowledge about the genus Absidia Tiegh. sensu lato regarding the traditional and current position within the order Mucorales (Mucoromycotina, Zygomycetes), nomenclatural changes and the taxonomical rearrangements of the prevalent species. Species grouped within Absidia possess some very promising industrial and medical application possibilities e.g., as mediators of biotransformations, producers of antimicrobial and wound healing stimulators. But some species are also causatives of severe and frequent fatal mucormycoses. Using traditional and modern methods of species determination had uncovered a trichotomous genera separation, namely Absidia sensu stricto, Lichtheimia, and Lentamyces, belonging to distinct families, the Absidiaceae, Lichtheimiaceae and “Lentamycetaceae” (NN), respectively. The existing medical and industrial aspects necessitate a fast and secure identification of a prominent species. Within this survey, morphological criteria and molecular markers were proposed for clear distinction of Absidia sensu lato. 19.1 The Genus Absidia: Current State of the Art – A Brief Survey The order Mucorales within the proposed subphylum Mucoromycotina (Hibbett et al. 2007) comprises ubiquitous soil fungi mainly living as saprobes, but also facultative parasites on other fungi or plants and opportunistic pathogens causing mucormycoses in human and animals (Ribes et al. 2000; Thirion-Delalande et al. 2005). One of these causative organisms is the genus Absidia causing rhino-cerebral K. Hoffmann Institute of Microbiology, School of Biology and Pharmacy, University of Jena, Neugasse, 25, 07743 Jena, Germany e-mail: Hoffmann.Kerstin@uni-jena.de Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_19, # Springer-Verlag Berlin Heidelberg 2010 439 440 K. Hoffmann mycoses, primary cutaneous, pulmonary or gastrointestinal lesions which occur sometimes in humans showing debilitating conditions like diabetes, burn wounds, immunosuppression, or traumata (Gonzalez et al. 2002; Ribes et al. 2000). Since opportunism of Absidia species is pronounced by growth maxima above body temperature, thermotolerant species enjoy physiological advantages over mesophilic ones. Not only due to its opportunism is the genus Absidia of recurrent interest but also due to some interesting biotechnological applications. For instance, microbial biotransformations are an important tool for pharmaceutical, economical purposes or even biodegradation of environmental pollutants (Chen et al. 2007; Demirci et al. 2004; Guiraud et al. 2008). To describe and classify species of the genus Absidia was therefore the main objective of several mycologists and taxonomists. The genus was described in 1876 by van Tieghem and augmented by reputable taxonomists like Bainier (1882, 1889), Hagem (1908), and Lendner (1907, 1908, 1924). Comprehensive morphological and physiological investigations of Absidia species were accomplished especially by Hesseltine, Ellis and Schipper (Hesseltine and Ellis 1961, 1964, 1966; Ellis and Hesseltine 1965, 1966; Schipper 1990). Molecular aspects have led to a renewal of species differentiation and were included in recent years, allowing for a reliable discrimination at species and genus level. Combined with differences in growth temperature as well as morphological distinctions, the genus Absidia was previously distinguished in Absidia sensu stricto encompassing the mesophilic species and in Lichtheimia comprising the thermotolerant species, which were summarized into a distinct family, the Mycocladaceae (orthographically erroneously pronounced as “Mycocladiaceae,” Hoffmann et al. 2007) and revised in the Lichtheimiaceae (Hoffmann et al. 2009b). For potential mycoparasitic species, a third genus, Lentamyces, was erected (Hoffmann and Voigt 2008). A detailed chronological summary of important accepted species and genera is given in Table 19.1 with family affiliation, type strain specification and MycoBank number (www.mycobank.org; Crous et al. 2004; Robert et al. 2005). 19.2 The Traditional Genus Absidia: Morphological Aspects Nomenclatural designation of Absidia (Etym.: absis, arcus) started as early as 1876 with the genus description by van Tieghem. He characterized this genus by (1) arcuated stolons with rhizoids, (2) rhizoids not opposite the sporangiophores, (3) sporangiophores arising from the elevated parts of the stolons, (4) apophysate and pyriform sporangia with deliquescent walls, and (5) zygospores surrounded by appendages originating from the suspensors (van Tieghem 1876). The assignment of species to the genus Absidia sensu lato is mainly based on traditional morphological criteria and growth parameters. In so doing, the last species identified so far is a thermotolerant variety of A. idahoensis (A. idahoensis var. thermophila, Chen and Zheng 1998). 19 Identification of the Genus Absidia 441 Table 19.1 Major species traditionally placed within Absidia sensu lato and currently accepted within the genera Absidia sensu stricto, Lichtheimia, and Lentamyces Year Species Type strain MycoBank no. Absidia TIEGH. 1876, MB20001, type species: A. repens TIEGH., family Absidiaceae MB81973 1876 A. repens TIEGH. CBS115583 (IT) MB223578 NRRL1315 (NT) MB351936 1889 A. caerulea BAINIER NRRL2797 (T) MB224063 1907 A. spinosa var. spinosa LENDN. 1908 A. cylindrospora HAGEM var. cylindrospora NRRL1317 (T) MB427391 NRRL1328 (T) MB221208 1908 A. glauca HAGEM NRRL2800 (T) MB252572 1930 A. heterospora Y. LING NRRL2793 (T) MB252285 1936 A. fusca LINNEM. 1958 A. spinosa var. azygospora BOEDIJN NRRL2841 (T) MB346488 NRRL2632 (T) MB292052 1959 A. cuneospora G.F. ORR & PLUNKETT 1961 A. cylindrospora var. rhizomorpha HESSELT. & J.J. ELLIS NRRL2771 (T) MB348992 NRRL2770 (T) MB325715 1962 A. pseudocylindrospora HESSELT. & J.J. ELLIS NRRL1807 (T) MB325709 1964 A. anomala HESSELT. & J.J. ELLIS NRRL3060 (T) MB353238 1964 A. cylindrospora var. nigra HESSELT. & J.J. ELLIS 1964 A. spinosa var. biappendiculata RALL & SOLHEIM NRRL3033 (T) MB348993 NRRL2968 (T) MB325710 1965 A. californica J.J. ELLIS & HESSELT. CBS697.68 (T) MB325712 1968 A. macrospora VÁNOVÁ Lichtheimia VUILL. 1903, MB20308, type species: L. corymbifera (COHN) VUILL., family Lichtheimiaceae MB508680 NRRL2981 (NT) MB416447 1903 L. corymbifera (COHN 1884) VUILL. NRRL1309 (NT) MB416448 1903 L. ramosa (ZOPF 1890) VUILL. 2009 L. hyalospora (SAITO 1906) K. HOFFM, G. WALTHER & K. VOIGT NRRL2916 (NT) MB512830 2009 L. ornata (A.K. Sarbhoy 1965) A. ALASTRUEY-IZQUIERDO & NRRL10293 (IT) – G. WALTHER – 2010 L. sphaerocystis A. ALASTRUEY-IZQUIERDO & G. WALTHER CBS420.70 (T) Lentamyces K. HOFFM. & K. VOIGT 2008, MB511979, type species: L. parricida (RENNER & MUSKAT ex HESSELT. & J.J. ELLIS) K. HOFFM. & K. VOIGT 2008 L. parricida (RENNER & MUSKAT ex HESSELT. & J.J. ELLIS NRRL2409 (T) MB511980 1964) K. HOFFM. & K. VOIGT NRRL2806 (T) MB511981 2008 L. zychae (HESSELT. & J.J. ELLIS 1966) K. HOFFM. & K. VOIGT NT Neotype; IT Isotype; T type; NRRL Northern Regional Research Laboratories, strain collection of the National Center of Agricultural Utilization Research Peoria, IL, USA; CBS Centraalbureau voor Schimmelcultures Utrecht, The Netherlands Although the first species described possess appendaged suspensors of their zygospores, several species lacking appendages were described in the following years, and due to further differences in morphology and physiology (especially sporangiophore branching, stolon, rhizoid and zygospore appearance, as well as growth temperature) ten different generic names were proposed over the years 1888–2008 in order to delimitate Absidia-like species from unequivocal designation to Absidia: Tieghemella BERL. & DE TONI (1888), Mycocladus BEAUVERIE (1900), Proabsidia VUILL (1903), Lichtheimia VUILL (1903), Pseudoabsidia [as “Pseudo-Absidia”] BAINIER (1903), Protoabsidia NAUMOV (1935), Gongronella RIBALDI (1952), Chlamydoabsidia HESSELT. & J.J. ELLIS (1966), Siepmannia KWAŚNA & NIRENBERG (2008a, b) and Lentamyces K. HOFFM. & K. VOIGT (2008). 442 K. Hoffmann The genera Absidia, Tieghemella, and Proabsidia are now considered as synonyms for species with zygospores surrounded by appendages from the suspensors whereas, Lichtheimia, Mycocladus, Pseudoabsidia, and Protoabsidia were assigned to Absidia species lacking such appendages (Hesseltine and Ellis 1964; Schipper 1990). The genus Gongronella, which is based on Absidia butleri LENDN., shows an apophysis, nonpyriform sporangia with reduced columellae, and zygospore suspensors devoid of appendages (Ribaldi 1952). Based on morphology, the genus Chlamydoabsidia is obviously nested within Absidia sensu stricto but developing unique multiseptate, pigmented aerial chlamydospores (Hesseltine and Ellis 1966). Most recently the species with warty exospore of their Mucor-like zygospores were separated in the genus Lentamyces, which was first invalidly (then correctly) described as Siepmannia the same year (Hoffmann and Voigt 2008; Kwaśna and Nirenberg 2008a, b). 19.3 The Impact of Molecular Data: The Genus Absidia sensu lato is Distinguished in at Least Three Nonrelated Genera Support of its morphological and physiological evidence by molecular based phylogenetic analyzes has led to a wide acceptance of the polyphyly of Absidia in recent years (Voigt et al. 1999; Voigt and Wöstemeyer 2001; O’Donnell et al. 2001; Kwaśna et al. 2006). This phylogenetic interpretation approximates a natural system contradicting the nomenclatural inflation of Absidia and its allied genera around the turn of the nineteenth century. In order to prove and support the relationship of the described species to a specific genus within this study, molecular data was analyzed in a multigene phylogeny and supplemented with morphological and physiological data (here especially parameters of growth temperature). An alignment of the combined ribosomal DNA sequences (18S rDNA, 28S rDNA) and nucleotide sequences coding for actin (act) and translation elongation factor 1 alpha (tef) was subjected to a Bayesian inference (Fig. 19.1). Since their morphology-based description, the genera Gongronella and Chlamydoabsidia are considered as independent Absidia-like genera, but they are still somehow related to Absidia morphologically and also on the molecular level, with Chlamydoabsidia nested within the genus and Gongronella closely related to the Absidia core group (Voigt et al. 1999; O’Donnell et al. 2001; Voigt and Wöstemeyer 2001; Hoffmann and Voigt 2008; Fig. 19.1). The classification currently accepted assigns Absidia-like taxa to three different genera and is supported by molecular phylogenies, morphological and physiological studies: (1) Absidia sensu stricto (syn. Tieghemella, Proabsidia), (2) Lentamyces and (3) Lichtheimia (syn. Mycocladus, Pseudoabsidia, Protoabsidia). 19 Identification of the Genus Absidia 443 Mortierella multidivaricata Mortierella verticillata Mortierella alpina Umbelopsis isabellina Umbelopsis nana Umbelopsis ramanniana Lentamyces parricida Lentamyces zychae Phycomyces blakesleeanus Spinellus fusiger Radiomyces spectabilis Saksenaea vasiformis Thermomucor indicae-seudaticae Rhizomucor miehei Rhizomucor pusillus Dichotomocladium elegans 93 Lichtheimia corymbifera Lichtheimia ramosa L. hyalospora L. hyalospora Rhizopus oryzae 93 Rhizopus stolonifer Mycotypha africana Mycotypha microspora Blakeslea trispora Choanephora cucurbitarum 99 Chaetocladium brefeldii Mucor racemosus Parasitella parasitica Gongronella butleri Hesseltinella vesiculosa Cunninghamella bertholletiae Cunninghamella echinulata Halteromyces radiatus Absidia spinosa Absidia psychrophilia Absidia repens 98 Absidia glauca Chlamydoabsidia padenii Absidia macrospora Absidia californica 0.1 substitutions / site Absidia caerulea Fig. 19.1 Phylogeny of inferred by Bayesian analysis from a combined analysis of aligned nucleotide sequences coding for actin, translation elongation factor 1alpha, small (18S) and large (28S) subunit rRNA (see Table 19.3). Posterior Probabilities (PP) are given above the branches with dots indicating 100% 444 K. Hoffmann 19.3.1 The Genus Absidia TIEGH. sensu stricto Deserves a Separate Family, the Absidiaceae V. ARX As already mentioned, the genus Absidia sensu stricto harbors mesophilic species with the optimum and maximum temperatures for growth around 30 C and 37 C respectively. Along with their common pyriform, multispored sporangia, they form columellae with species specific apical projections (Hoffmann et al. 2007). Furthermore, the asexually produced sporangiospores are species-specifically globose to cylindrical. As outlined in detail by Schipper (1990), Absidia could be divided into several groups distinguishable by their spores. This distinction is of taxonomical importance, since it is also present in molecular phylogenetic analyzes. (Kwaśna et al. 2006; Hoffmann et al. 2007, 2009a; Hoffmann and Voigt 2008). Within these analyzes three well supported clades are obvious: A. caerulea, A. californica, A. glauca, and A. macrospora with globose sporangiospores (temperature for growth and sporulation at 15–30 C, no growth at 34–37 C). A. anomala, A. cylindrospora, A. pseudocylindrospora, A. psychrophilia, A. repens, and A. spinosa possess oval to cylindrical spores (growth and sporulation at 15–34 C, no growth at 30–37 C), and A. cuneospora shows conical shaped spores and represents an intermediate species positioned between the other two groups (Schipper 1990; Hoffmann et al. 2007; missing data for A. cuneospora in Fig. 19.1). Species within these clades, especially A. caerulea and A. glauca, offer some pharmaceutical important properties as biomimetic models to test the metabolism of xenobiotics like steroids and saponins in mammalia (Brezezowska et al. 1996; Huszcza and Dmochowska-Gladysz 2003; Chen et al. 2007). Both species are also potentially chitosan producers, useable in the food processing industry as well as medical applications since chitosan seems to possess antimicrobial and woundhealing stimulating activities (Abdel-Fattah et al. 1984; Muzzarelli et al. 1994; Rungsardthong et al. 2006; Dai et al. 2009). Some species were also extensively studied for applications in environmental detoxifications like A. cylindrospora and A. fusca, which are able to degrade polycyclic aromatic hydrocarbons (Guiraud et al. 2008). All species belonging to the genus Absidia sensu stricto could now be summarized in the family Absidiaceae, which was erected by von Arx (1982). Although von Arx included fourteen genera within the Absidiaceae, only six are accepted here, namely Absidia, Chlamydoabsidia, Cunninghamella, Gongronella, Halteromyces, and Hesseltinella. Even though, there are only a few studies including all affected genera in one analysis, the close relationships between some of these genera was repeatedly demonstrated. Based on 18S and 28S rDNA sequences, Absidia, Chlamydoabsidia and Cunninghamella are well supported sister taxa (100% BS, Voigt et al. 1999). Based on the combined coding sequences of act and tef, Absidia, Chlamydoabsidia, Cunninghamella, Hesseltinella and Halteromyces are in close vicinity (without high support values, Voigt and Wöstemeyer 2001) but high support values for this clade (additionally including Gongronella, 100% BS) are presented by O’Donnell et al. 2001, in an analysis of combined tef and rDNA 19 Identification of the Genus Absidia 445 sequences. Except of Chlamydoabsidia (which is always nested within Absidia), all genera are slightly distinct from the core Absidia, but are related within one monophyletic clade (Fig. 19.1, 100% posterior probability (PP)). On the morphological level, the main distinctions could be found in the shape of the sporangia (e.g., pyriform for Absidia, dumbbell-shaped for Halteromyces), number of spores produced within each sporangia (e.g., few or uni-spored [Cunninghamella] or multispored [Halteromyces, Absidia]), and zygospore appearance (with [Absidia] or without appendages on the suspensors [Cunninghamella, Gongronella]). All of these characteristics seem not to be systematically important in family description, a finding which was also proven for other families within the order Mucorales (Voigt et al. 2009), and in the past, these morphological criteria have led to systematic classifications, not sustainable in the era of molecular phylogenetics (e.g., Absidiaceae comprised fourteen genera from which only six are phylogenetically related; compare von Arx 1982 and O’Donnell et al. 2001; Voigt and Wöstemeyer 2001). Although Absidiaceae is treated as a synonym of the highly polyphyletic family Mucoraceae (Kirk et al. 2008), in accordance to Voigt et al. (2009) a restoration of the monophyletic Absidiaceae is highly recommended here. 19.3.2 The Genus Lichtheimia (COHN) VUILL. and Its Family Lichtheimiaceae K. HOFFM., G. WALTHER and K. VOIGT Lichtheimia corymbifera (formerly: Absidia corymbifera) is a common causative agent of mucormycoses and was described as Mucor corymbifer by Cohn 1884. Mycoses caused by other species belonging to this genus were not reported yet, which is, in all likelihood, a problem of proper identification as well as taxonomical knowledge. Morphological and physiological separation between the species of Lichtheimia depends on differences in the growth temperature, in the giant cell and partly in the sporangiospore morphology (Hoffmann et al. 2007, 2009a; AlastrueyIzquierdo et al. 2010), and with optima for growth around 37 C all species should be capable to colonize endothermic organisms under appropriate conditions. With the growing awareness for the need to differentiate between single species in fungal infections, it was essential to discriminate Lichtheimia from Absidia species, because species of the mesophilic genus Absidia are known to be harmless and do not cause systemic infections in human and warm-blooded animals. Furthermore, the assignment of all known former thermotolerant Absidia species to the genus Lichtheimia is essential for the establishment of a natural monophylum-based system. Lichtheimia is distinguished from Absidia not only on the molecular level, but also on the basis of morphological and growth-physiological differences. Since comprehensive analyses including as many genera as possible, thermotolerant species were clearly distinct from mesophilic Absidia species (Voigt et al. 1999, 2009; O’Donnell et al. 2001; Voigt and Wöstemeyer 2001; Hoffmann et al. 2007; Hoffmann and Voigt 2008). Due to zygospores lacking appendaged suspensors and growth temperatures with optima and maxima above 37 C, a separation of these 446 K. Hoffmann species within the subgenus Mycocladus was proposed by Hesseltine and Ellis (1964) and was accepted by Schipper (1990). This subgenus was based on the type taxon M. verticillatus BEAUVERIE (Beauverie 1900). Mycocladus was also described as independent genus (Mirza et al. 1979; Vánová 1991). Based on physiological, phylogenetic and morphological analyses the (sub-)genus Mycocladus was reclassified in the new family Mycocladaceae [as “Mycocladiaceae”], K. HOFFM., DISCHER & K. VOIGT emphasizing its phylogenetic distinctness from species designated to Absidia (Hoffmann et al. 2007). But in the course of a critical reexamination of all available literature references, especially the years around the turn of the century (1882–1904), the correct assignment of Mycocladus verticillatus as type taxon for nonappendaged zygospore suspensors was rejected. Furthermore, on the morphological level Mycocladus verticillatus seems to be a coculture of at least two different species (Beauverie 1900; Hoffmann et al. 2009b). The first rough-walled zygospores with unadorned suspensors and equatorial ridges within the genus Absidia were described in 1903 by Bainier for the species Pseudoabsidia vulgaris [as “Pseudo-Absidia”], based on Absidia dubia BAINIER 1882 (Bainier 1882, 1903). His error of assigning a wrong epithet was corrected the same year by Sydow, proposing Pseudabsidia dubia (BAINIER) SYDOW (Sydow 1903). The species was later transferred to Lichtheimia, a genus named in honor of the mycologist Lichtheim, professor at the University of Bern, Switzerland (Vuillemin 1903, 1904). Within the same publication issue of the Bulletin de Socie´te´ Mycologique de France where Bainier described Pseudoabsidia (page 155), Vuillemin proposed the genus Lichtheimia (page 126). He recognized Mucor corymbifer COHN as type of Lichtheimia. The appearance of nonappendaged and naked zygospores in L. corymbifera was described separately (Vuillemin 1904). The original description of Lichtheimia VUILL. and its type L. corymbifera represents the new family Lichtheimiaceae K. HOFFM., G. WALTHER & K. VOIGT which harbors the species L. corymbifera (COHN) VUILL., L. ramosa (ZOPF) VUILL., L. ornata (A.K. SARBHOY) A. ALASTRUEY-IZQUIERDO & G. WALTHER, L. hyalospora (SAITO) K. HOFFM., WALTHER & K. VOIGT and L. sphaerocystis A. ALASTRUEY-IZQUIERDO & G. WALTHER. 19.3.3 The Genus Lentamyces and a Required Excursus to Siepmannia: Species Boundaries The genus Lentamyces K. HOFFM. & K. VOIGT comprises homothallic and potentially mycoparasitic species producing warty zygospores lacking appendaged suspensors (Hoffmann and Voigt 2008). This genus is composed of currently two species with the potentially mycoparasitic L. zychae and the mycoparasitic L. parricida. L. zychae was originally described to be mycoparasitic on other Mucorales (Zycha 1935), but could never be confirmed as causative agent of mycoparasitic reactions on other mucoralean hosts in recent confrontation experiments using the type strain NRRL2806 (Table 19.1; Hoffmann and Voigt 2008). But for L. parricida a wide 19 Identification of the Genus Absidia 447 variety of potential mucoralean hosts are experimentally confirmed (Schipper 1990; Hoffmann and Voigt 2008). Apart from the varying potential for mycoparasitic abilities, both species show morphological and physiological characteristics of mycoparasitic fungi such as slowly developing colonies with thin mycelia, and abundant sucker-like branches in the substrate mycelium of single cultures but vigorously growing hyphae in cocultures with other mucoralean fungi. L. zychae and L. parricida form a monophyletic group (100% PP) which is closely related to the Phycomycetaceae (Voigt et al. 2009). The Lentamyces clade may represent the next to the last family, the putative “Lentamycetaceae,” at the base of the order Mucorales (Fig. 19.1). Their homothallic and facultative parasitic nature turns this basal lineage of mucoralean fungi into an interesting subject to study the evolution of fungal communication directly at the level of radiation and diversification of mucoralean fungi (Hoffmann et al. 2007; Hoffmann and Voigt 2008). In 2008 both species were classified into the genus Siepmannia as S. parricida and S. zychae. The genus was amended by two new species, S. pineti and S. lariceti. The authors described S. pineti as morphologically similar to Fennellomyces linderi and S. lariceti similar to Mucor circinelloides but with micromorphological features of different size (Kwaśna and Nirenberg 2008a). Nevertheless, judging by the presented descriptions and photographs, S. pineti seems more like a species of the genus Circinella: developing structures typical for Circinella like circinate sporangiophores terminating in globose to dorsiventrally flattened sporangia with a persistent wall. F. linderi would occasionally develop a subsporangial vesicle, not described for Siepmannia. S. lariceti seems with all likelihood representing a species of Mucor, as stated by Kwaśna and Nirenberg. On the contrary, neither L. parricida nor L. zychae produce circinate nor sympodially branched sporangiophores as described for Siepmannia. The sporangia are always globose to subpyriform, apophysate surrounded with a deliquescent sporangial wall (Hesseltine and Ellis 1964, 1966). Although species of Siepmannia are described to be mycotrophic, no hint for potential mycoparasitism was reported (Kwaśna and Nirenberg 2008a). Parasitic interactions, e.g., between L. parricida and one of its host Zygorhynchus moelleri, was successfully demonstrated by Hoffmann and Voigt 2008. A typical morphological feature of Lentamyces is the abundant sucker-like substrate mycelium and the homothallic formed globose zygospores with a warty exospore. No such typical mycelium was pictured for S. lariceti and S. pineti. The zygospores of S. lariceti are globose to ellipsoid and not as large and warty as those of Lentamyces. The only criteria which have Siepmannia and Lentamyces in common are of physiological nature, e.g., a low maximum temperature for growth and a restricted growth, which is stimulated in the presence of other fungi (potential hosts in the case of Lentamyces) (Kwaśna and Nirenberg 2008a; Hesseltine and Ellis 1964, 1966; Hoffmann and Voigt 2008). However, physiological criteria may vary among isolates and species and their relevance as synapomorphic characters is limited (Voigt et al. 2009). In a further step, the authors generated species-specific restriction patterns of digested ITS sequences. They used the enzymes AluI, HhaI, DdeI, HaeIII, HincII, HinfI, HpaII, Sau3AI, and TaqI, which clearly differentiated S. pineti 448 K. Hoffmann from S. lariceti and F. linderi as well as S. lariceti from M. circinelloides (Kwaśna and Nirenberg 2008a) Nevertheless, they did not say anything about the restriction patterns of L. parricida and L. zychae. Repeating these restriction experiments with all species of Lentamyces and Siepmannia in theoretical in silico RFLP analyses using BioEdit v.7.0.9.0 (Hall 1999) resulted in highly variable restriction fragment length polymorphisms among the strains (Table 19.2). Neither Siepmannia nor Lentamyces show similar restriction sites to each other or to Mucor circinelloides or Fennellomyces linderi. Based on the present characteristics, it could not be stated whether Lentamyces and Siepmannia share a common generic designation or represent distinct taxa. The molecular phylogenetic analysis presented by Kwaśna and Nirenberg (2008a) was done on ITS1-5.8S rDNA-ITS2 sequences and lacks reference and type species. The fusion of both genera within Siepmannia neglects obvious morphological and molecular distinctions and does not doubtlessly justify a combination of the genus Siepmannia with L. zychae and L. parricida. The relationship between Siepmannia and Lentamyces has to be proven by further investigations Table 19.2 Theoretical in silico RFLP analyses of species belonging to the genera Lentamyces and Siepmannia as well as Mucor circinelloides and Fennellomyces linderi. Bold indicated restriction sites are similar in Siepmannia but are not present in Lentamyces, M. circinelloides or F. linderi Species and Length of Restriction sites for different enzymes GenBank ITS1AluI HhaI DdeI HaeIII HincII HinfI HpaII Sau3AI TaqI accession 5.8S-ITS2 number [bp] S. pineti 630 241 – – 473 – 339 151 25 168 AJ748134 249 347 381 272 291 499 557 509 S. lariceti 402; first 547 – – 466 – 340 – 272 291 178 bp AJ748857 are missing 348 L. parricida 545 – – – 406 – 267 309 198 90 AY944884 275 472 100 217 L. zychae 594 444 – – – – 10 331 222 119 EF030529 289 512 241 297 544 L. zychae 594 444 – – 422 – 10 331 222 119 AJ968561 289 343 241 297 512 544 560 M. circinelloides 556 129 337 78 147 122 308 237 256 AJ878535 142 457 316 314 533 F. linderi 669 467 34 354 283 72 AJ878536 211 601 302 482 656 360 19 Identification of the Genus Absidia 449 in the context of extended multilocus phylogenetic analyzes on a taxon set which comprehensively represents the Mucorales as a whole as demonstrated by O’Donnell et al. (2001) and Voigt and Wöstemeyer (2001). Furthermore, the first descriptions of the genus Siepmannia as well as the new combinations of S. zychae and S. parricida were not in accordance with the articles 37.1, 33.4 and 43.1 of the ICBN (McNeill et al. 2006). Although a validation was made the same year (Kwaśna and Nirenberg 2008b), the genus Lentamyces for the species L. parricida and L. zychae was validly published before (Hoffmann and Voigt 2008). The final phylogenetic position of the genus Siepmannia and its relation to Lentamyces requires further investigations; even a combination in one family is possible. 19.4 Morphological and Molecular Differentiation of Affected Species Belonging to Absidia sensu lato Although the traditional approach of comparative morphology for species differentiation is quite easy to perform, some if not most of the existing descriptions keys are now outdated. The more the data gained, the more changes in taxonomical designations were performed as a consequence of varying interrelationships. The increase of importance of molecular data combined with a broad range of analytical tools allows improvements in statistically supported phylogenetic relationships at all taxonomical levels. Natural relationships among species could not predict and be fully resolved on the basis of morphology alone but by comparative molecular phylogenetics as well, and therefore, in order to assign an organism to its natural affiliation, morphology should be supplemented with additional and independent data. In the following a synoptic key to the affected genera and species of Absidia sensu lato is given, which extends the detailed descriptions given by Ellis and Hesseltine (1965, 1966), Hesseltine and Ellis (1961, 1964, 1966), Zycha et al. (1969), Schipper (1990) and Alastruey-Izquierdo et al. (2010). In addition short signature consensus sequences and PCR-restriction fragment length polymorphisms of the ribosomal ITS1-5.8S rDNA-ITS2 region are proposed for genus separation. 19.5 Synoptic key to Genera and Species 1a. thermotolerant, temperature optimum and maximum above 34 C, growth above 37 C, rapidly growing, sporangiophores often without subsporangial septum, non parasitic on other Mucorales, zygospores without appendaged suspensors ........................................................................... genus Lichtheimia 1b. not thermotolerant, temperature maximum above 30 C, no growth above 37 C, rapidly growing, sporangiophores with subsporangial septum, non parasitic on other Mucorales, zygospores with appendaged suspensors ....... genus Absidia 450 K. Hoffmann 1c. not thermotolerant, temperature maximum below 30 C, slowly growing, sporangiophores with subsporangial septum, potentially parasitic on other Mucorales, homothallic, warty zygospores without appendaged suspensors .............. ............................................................................................... genus Lentamyces 19.5.1 The Genus Lichtheimia (VUILL. 1903; Lichtheimiaceae K. HOFFM., G. WALTHER & K. VOIGT 2009): This Key Is Originally Published in Alastruey-Izquierdo et al. (2010) 1a. Sporangia dark brown or dark grey to black; colony diameter after 72 h at 43 C < 2 mm; mature sporangiospores rough and/or > 6.5 mm in their longest extension ................................................................................................. 2 1b. Sporangia light brownish grey; colony diameter after 72 h at 43 C > 14 mm; mature sporangiospores smooth and < 6.5 mm in their longest extension .....3 2a. Giant-cells consistently globose, 60–150 mm in diameter ....L. sphaerocystis 2b. Giant-cells (if present) more hypha-like, irregularly swollen, simple to strongly branched, never consistently globose ...................... L. hyalospora[1] 2ba. Mature sporangiospores small (< 5.5 mm), rough, and brownish .................. ......................................................... small-spored variants of L. hyalospora[1] 2bb. Mature sporangiospores larger (on the majority > 5.5 mm), smooth or rough, hyaline or brownish ........................large-spored variants of L. hyalospora[1] 3a. Colony diameter after 72 h at 43 C > 40 mm, spores ellipsoidal to cylindrical or subglobose to broadly ellipsoidal ................................................L. ramosa 3b. Colony diameter after 72 h at 43 C < 27 mm, spores never consistently ellipsoidal to cylindrical ................................................................................ 4 4a. Densely branched giant-cells, 380–760 ( 900)  320–660 ( 770) mm, present in 2-week-old YEA cultures ....................................................... L. ornata 4b. Giant-cells absent from 2-week-old YEA cultures .................L. corymbifera [1] L. hyalospora is now a synonym of L. blakesleeana (Alastruey-Izquierdo et al. 2010). L. hyalospora was originally seperated from L. blakesleeana by the formation of larger and unusual hyaline mitospores. A prospective new separation of both species in varieties or formae could not be excluded. 19.5.2 The Genus Absidia (Tiegh. 1876; Absidiaceae Arx 1982) 1a. all sporangia pyriform, apophysate, with prominent columella, columellae often with apical projection, without chlamydospores ................ 2 (Absidia) 1b. dark-colored chlamydospores within aerial hyphae ........................................ ................................................................................. Chlamydoabsidia padenii 19 Identification of the Genus Absidia 451 2a. sporangiophores circinate and arising individually, sporangia mutant ....... ........................................................................A. reflexa (uncertain species) 2b. sporangiophores not bent ........................................................................... 3 3a. sporangiospores smooth ............................................................................. 4 3b. sporangiospores roughened to echinulate ..... A. scabra (uncertain species) 4a. sporangiospores spherical, heterothallic .................................................... 5 4b. sporangiospores spherical, homothallic ........A. septata (uncertain species) 4c. sporangiospores throughout not spherical, heterothallic or homothallic ......8 5a. all sporangiospores spherical 2.5–5.5 mm ................................................. 6 5b. all sporangiospores spherical up to 8.9 mm ......................... A. macrospora 6a. young mycelium violet ............................................................. A. caerulea 6b. young mycelium of different color ............................................................ 7 7a. sporangiophores single or in whorls, often two; young mycelium strain specifically colored (white, green, brown) .................................. A. glauca 7b. sporangiophores single or in abundant whorls, mostly more than two; young mycelium green to grey .........................................................A. californica 8a. sporangiospores more or less cylindrical, species homothallic ................. 9 8b. sporangiospores diverse (oval, cylindrical, globose, conical, irregular), species heterothallic ................................................................................. 10 9a. young cultures violet or reddish ............................................... A. anomala 9b. young cultures white, never violet or reddish, if suspensors typically unequal in size then appendages originate from the larger one ..................... .................................................................................... A. spinosa var. spinosa 9c. zygospore suspensors equal in size, appendages originating from both suspensors ................................................A. spinosa var. biappendiculata 9d. similar to variety spinosa but abundant azygospores with up to 3 spores on one suspensor ...................................................A. spinosa var. azygospora 10a. abundant secondary sporangia in older cultures, sporangiospores oval to short cylindrical ............................................................................A. repens 10b. no secondary sporangia, sporangiospores conical or clearly cylindrical ....11 11a. sporangiospores conspicuously conical ................................A. cuneospora 11b. sporangiospores diverse, irregularly shaped, globose and cylindricalellipsoidal, columellae without distinct apical projections .... A. heterospora 11c. sporangiospores regularly cylindrical ...................................................... 12 12a. young colonies on PDA white, older colonies greyish-brown, sporangiophores in whorls (1–4), zygospores with unequal suspensors ..................................... .................................................................A. cylindrospora var. cylindrospora 12aa. appears similar to A. pseudocylindrospora but older colonies on PDA medium of dark brownish grey color, no mating with A. cylindrospora .... .......................................................................... A. cylindrospora var. nigra 12ab. colonies on PDA light greyish brown; on Czapek-agar forming rhizomorph-like hyphae, no mating with A. cylindrospora ................................ ..............................................................A. cylindrospora var. rhizomorpha 12b. similar to A. cylindrospora, but older colonies blackish brown pigmented, sporangiophores in whorls up to six ...............................................A. fusca 452 K. Hoffmann 12c. similar to A. cylindrospora, young colonies grey, zygospores with equal suspensors ............................................................. A. pseudocylindrospora 12d. young colonies brownish, temperature optimum between 15–20 C, no growth at 30 C .................................................................. A. psychrophilia 19.5.3 The Genus Lentamyces (K. HOFFM. and K. VOIGT 2008) 1a on MEX mycelium not higher than 3 mm, thin, slow growing, brownish, facultative parasitic on other Mucorales, abundant zygospores, sporangiospores cylindrical 1.6–2.5 mm  1.9–3.3 mm ..................................................L. parricida 1b not parasitic on other Mucorales, rare zygospores, sporangiospores cylindrical 1.2–2.2 mm  1.6–3.3 mm ........................................................................L. zychae 19.6 Some Remarks to Uncertain Species Among the literature species epithets exist which are only once encountered and described. The species are not available in any culture collection and, if there was no mistake in the observations described, could be presumed as lost. Therefore, A. scabra, A. septata and A. reflexa still remain within the genus Absidia until they are rediscovered because the passed down morphological descriptions match those of Absidia. Except for the roughened to echinulate sporangiospores and missing septae beneath the sporangium, A. scabra COCCONI resembles the characteristics of A. caerulea (Cocconi 1899; Ellis and Hesseltine 1965). The description of A. septata TIEGH. is also nearly identical with those of A. caerulea but differs in the presence of homothallically formed zygospores which were definitively illustrated (van Tieghem 1876; Ellis and Hesseltine 1965). A. reflexa TIEGH. described by van Tieghem (1876) shows obvious circinate sporangiophores which is not typical for Absidia (Ellis and Hesseltine 1965). Until a rediscovery and without a profound phylogenetic study, a clear designation to any species or genera remains uncertain. Because they lacked sufficient diagnoses or absence in appropriate strain collections the following taxa were not considered here: A. aegyptiacum SATORY, MEYER & TAWFIK 1939; A. capillata TIEGH. 1876; A. clavata B.S. MEHROTRA & NAND 1967; A. fassatiae VÁNOVÁ 1971; A. griseola H. NAGAN. & HIRAHARA 1970; A. inflata J.H. MIRZA, S.M. KHAN, S. BEGUM & SHAGUFTA 1979; A. narayanai SUBRAHAMANYAM 1990; A. robusta RACIBORSKI 1899; A. tuneta RENNER & MUSKAT 1958; A. ushtrina S.C. ARARWAL 1974. 19.7 Molecular Key to the Genera The augmentation of generally easy-to-access morphological data with molecular characteristics will support the current concepts of species relationships to specific genera. Sequences of the nuclear internal transcribed spacer (ITS) region already 19 Identification of the Genus Absidia 5` GGCACRGTTGTTTCAGTATC 3` 5` CCGGWGRGKACGCCTG 3` 5` GGTACGYCTGTTTCAGTATCATT 3` 18S rDNA 453 Lichtheimia Lentamyces Absidia ITS1 5.8S rDNA 5` GGAAGGATCATTACTGAGAGG 3` 5` CGGAAGGATCATTAMTGTTTWTG 3` 5` TGCGGAAGGATCATTARAAATG 3` Lichtheimia Lentamyces Absidia 5` GATCTGAAATCAACTGAGAYYAC 3` 5` GYCTGAAATCAGGTGGGATTAC 3` 5` TTGATCTGAAATCAGRYGGGA 3` ITS2 28S rDNA 5` ATGGATCTCTTGGTTCTCGCA 3` 5` CGGATCTCTTGGTTCTCGCA 3` 5` TGGATCTCTCGGCTTTCGTATC 3` Fig. 19.2 Schematic illustration of the nuclear ribosomal DNA cluster including internal transcribed spacer regions (ITS) 1 and 2 with short signature sequences to differentiate between the genera Absidia sensu stricto, Lentamyces and Lichtheimia. Signature sequences are located at the 18S rDNA – ITS1 boundary, at the 50 and 30 end of the 5.8S rDNA, and at the 50 end of the 28S rDNA. Positions of the signature sequences could slightly vary …C TCGAG… Lentamyces …GAGCT C… Lichtheimia ramosa AY944897 18S rDNA ITS1 Absidia 5.8S rDNA ITS2 28S rDNA …TCG CGA… …AGC GCT… Fig. 19.3 Schematic illustration of the nuclear ribosomal DNA cluster including internal transcribed spacer regions (ITS) 1 and 2 with unique restriction sites useful for the discrimination of the genera Absidia sensu stricto, Lentamyces and Lichtheimia. Species in the genus Absidia sensu stricto possess a unique restriction site for the enzyme Bsp68I (NruI, recognizing the motif TCGCGA) near the 50 end of the 5.8S rDNA, cutting ITS sequences (between 500–700 bp) in two fragments nearly about the same size. In the middle of the ITS1 sequence (around position 100) is a XhoI restriction site located (recognizing the motif CTCGAG), unique for the genus Lentamyces. With the exception AY944897 both sequence motifs occur not within sequences of the genus Lichtheimia proved to be useful for species designation and phylogenetic analyzes concerning members of the genera Absidia, Lentamyces and Lichtheimia (Machouart et al. 2006; Schwarz et al. 2006; Hoffmann et al. 2007, 2009a). Several ITS sequences available from GenBank (www.ncbi.nlm.nih.gov; listed in the methodical Sect. 9.3.) were analyzed for short sequence fragments useful to differentiate between genera (Fig. 19.2). Signature sequences are located at the nuclear 18S ribosomal DNA – ITS1 boundary, at the 50 and at the 30 end of the 5.8S rDNA, and at the 50 end of the 28S rDNA. Positions of the signature sequences could slightly vary. Furthermore, restriction site analyzes of the sequenced PCR products revealed unique sites potentially useful for differentiation of the genera (Fig. 19.3). Within Absidia sensu stricto the enzyme Bsp68I (TCG CGA) cuts the total ITS sequences within the 5.8S rDNA region in two fragments nearly about the same size. There is no Bsp68I restriction site within the analyzed sequences of species belonging to the 454 K. Hoffmann genera Lentamyces or Lichtheimia (or Siepmannia). A unique recognition site for the enzyme XhoI is present within Lentamyces (but not within Siepmannia), nearly half cutting ITS1 (around position 100 of the total ITS1-5.8S rDNA-ITS2 sequence). Absidia and Lichtheimia show no XhoI restriction site with one exception. The ITS1 region of Lichtheimia ramosa (AY944897) is restricted by XhoI after two-thirds (Fig. 19.3). 19.8 Combining Morphological and Molecular Characters: An Example Bearing in mind that the name of a type strain is intrinsically tied to its morphological features, a molecular-based identification should not neglect a profound check on morphology. One example is illustrated by Absidia repens. This species is characterized by the presence of uniquely formed secondary sporangia beneath typical Absidia-like sporangia. A. repens is one of the few species with available sequences from different geographical populations. In Fig. 19.4. an alignment of three isolates is displayed. Two isolates are from America (A) and the isotype is from Europe (E). Both American isolates are similar to each other with 95% identity but differ considerably from the European isolate (52% identity). As already outlined in detail by Hoffmann et al. 2009a, information on the geographical origin could solve differences within isolates caused by geographically-based species deviation and will eventually support the erection of new formae, varieties or even species. Such a cryptic species with no obvious morphological differences is A. repens, but a separation in two distinct species requires a more profound investigation of the already described isolates. A synonym of A. repens, namely A. japonica, was isolated in Japan and may easily represent a distinct cryptic species. 19.9 Methodological Section 19.9.1 Media for Cultivation Media for cultivation, mentioned in the description keys were PDA (see DSMZ medium 129), Czapek-Agar (see DSMZ medium 130) and MEX (3% malt extract supplemented with 0.5% yeast extract); YEA (yeast extract, Difco, Alphen a/d Rijn, The Netherlands); MEA (malt extract, Difco). 19.9.2 Methods for Strain Maintenance, Cultivation, and Sequence Analyses Methods for strain maintenance, cultivation, and sequence analyses were previously described by Hoffmann et al. (2007) and Hoffmann and Voigt (2008). 19 Identification of the Genus Absidia 455 CBS101.32(A)1 -------------------------------------------------- 50 KAS3611 (A) 1 -------------------------------------------------- 50 CBS115583(E)1 GAAATGCTGGGAAGCCTCCGGGAGGACCTAACTTTTTTCTACTGGTCCCT 50 CBS101.32 KAS3611 CBS115583 51 -------------------------------------------------- 100 51 -------------------------------------------------- 100 51 TGTTTTTTTAGGGGGTTGCTTGGGAAGGGATTCGTTTCTTCCCTTGATGT 100 CBS101.32 101 -----AAAATGCGGCCGGTTCTCTTTCGGGAGGATTGGTCAACAGATTTA 150 KAS3611 101 -----AAAATGCGGCTGGCTCTCTTT--GGAGGGTTGGTCAACAGATTTA 150 CBS115583 101 TTGGGGGAATTTTATTATTCCCCCTTCATGGGAAAGTTTTACTACTTTCC 150 CBS101.32 151 ATTCTGTGCACTGTTTTTAATTGGGGGTTTTCTTGAAAAAGGGAGCCTCC 200 KAS3611 151 ATTCTGTGCACTGTTTTTAATTGGGGGTTTTCTTGAAAAAGGGAGCCTCC 200 CBS115583 151 CCTTCTCCCACCCTGGGTAAA-GCCCTTTTTCCTTTGGGAGAATCCGGTT 200 CBS101.32 201 TGCCCTGG-GTATTGCTCTTTTTCCTTTGGGAAGAAATCAGCTTGCCCTA 250 KAS3611 201 TGCCCTGG-GTATTGCTCTTTTTCCTTTGGGAAGAAATCAGCTTGCCCTA 250 CBS115583 201 TGCCCAGTTGAATTCCCCTTCTTTCATAGGGGGGGGG-----TTTTCAAG 250 CBS101.32 251 TTAATATACTATTCTGACTGAACTAAAACAGAAAATTGTTTAACACATAA 300 KAS3611 251 TTAATATACTATTCTGACTGAACTAAAACAGAAAATTGTTTAACATATAA 300 CBS115583 251 TTTATATACTATTTTGACTGAACTAAA-CAGAAA-TTGTTTAACACTTAA 300 CBS101.32 301 ACAACTTTCAGCAATGGATCTCTCGGCTTTCGTATCGATGAAGAACGCAG 350 KAS3611 301 ACAACTTTCAGCAATGGATCTCTCGGCTTTCGTATCGATGAAGAACGCAG 350 CBS115583 301 ACAACTTTCAGCAATGGATCTCTCGGCTTTCGTATCGATGAAGAACGCAG 350 CBS101.32 351 CAAATCGCGATATGTAGTGTGATCTGCCTATAGTGAATCATCAAATCTTT 400 KAS3611 351 CAAATCGCGATATGTAGTGTGATCTGCCTATAGTGAATCATCAAATCTTT 400 CBS115583 351 CAAATCGCGATATGTAGTGTGATCTGCCTATAGTGAATCATCAAATCTTT 400 CBS101.32 401 GAACGCATCTTGCACCCTTGGGTATTCCTGAGGGTACGCCTGTTTCAGTA 450 KAS3611 401 GAACGCATCTTGCACCCTTGGGTATTCCTGAGGGTACGCCTGTTTCAGTA 450 CBS115583 401 GAACGCATCTTGCACCCTTGGGTATTCCTGAGGGTACCCCTGTTTCAGTA 450 CBS101.32 451 TCATTTTAACTTCATCTCCTTTCGAGGGGTTTG----------AAAAAAT 500 KAS3611 451 TCATTTTAACTTCATCTCCTTTCGAGGG-TTTG----------AAAAAAT 500 CBS115583 451 TCATTTTTATCTTCTTTCCCGTCCTTAGGTTGGTGGGGAAGGGGAAAAAT 500 CBS101.32 501 CACTACTGGCCATTGAGTACCTTTGT------GTATTTCTCGGCTGAAAT 550 KAS3611 501 CACTACTGGCCATTGAGTACTTTATT------GTGCTTCTCGGCTGAAAT 550 CBS115583 501 CAC-ACTCGCC-TAGAGTACTAGTTTAGACCGGTGCTTCTTGGCTGAAAT 550 CBS101.32 601 AATCT-TATGGTTTCCCTTATGACTGGGGGCAATTACCCTTTGGTAGAAT 600 KAS3611 601 AATTT-TATGGTTTCCCTTATGACTGGGGGCAATTACCCTTTGGTAGAAT 600 CBS115583 601 TATTGATACAGTTTTCCCTTTGACTTTAAAGGGGTACCCTTTGGTAGCCT 600 CBS101.32 651 TTATTTTTTACAAAAGAAAAAATTGAAGCCAGTCTAGAAGCTATACCGTC 650 KAS3611 651 TCATTTTTTACAAAAGAAACAATTGAAGCCAGTCTAGAAGTCATACCTTC 650 CBS115583 651 TTC----------------------------------------------- 650 CBS101.32 701 GAAAGACAACCCCAAAAA 718 KAS3611 701 --------ACCCCAAAAA 718 CBS115583 701 ------------------ 718 Fig. 19.4 Alignment of sequenced ITS1-5.8S rDNA-ITS2 regions from three different isolates of Absidia repens. CBS101.32 and KAS3611 were isolated in America (A). The isotype of the species is of European origin (E). The European isolate differs primarily by a major indel (insertion/deletion) within ITS1 as well as unambiguously aligned regions within ITS1 and ITS2 of the American strains 456 K. Hoffmann Bayesian inference was done using MrBayes v3.1.2 (Huelsenbeck and Ronquist 2001; Ronquist and Huelsenbeck 2003). The alignment consists of combined sequences of partial 18S rDNA (1202 characters), partial 28S rDNA (389 characters), partial sequences coding for act (807 characters) and tef (1092 characters). Table 19.3 Species and GenBank accession numbers studied in the Bayesian inference (Fig. 19.1) Species GenBank accession numbers act tef 18S rDNA 28S rDNA Absidia caerulea EU736223 EU736245 EU736272 EU736299 A. californica EU736224 EU736247 EU736274 EU736301 A. glauca EU736225 EU736248 EU736275 EU736302 A. macrospora AY944760 EU736249 EU736276 EU736303 A. psychrophilia AY944762 EU736252 EU736279 EU736306 A. repens AJ287136 AF157228 AF113410 AF113448 A. spinosa EU736227 EU736253 EU736280 EU736307 Blakeslea trispora AJ287143 AF157235 AF157124 AF157178 Chaetocladium brefeldii AJ287144 AF157236 AF157125 AF157179 Chlamydoabsidia padenii AJ287146 AF157238 AF113415 AF113453 Choanephora cucurbitarum AJ287147 AF157239 AF157127 AF157181 Cunninghamella bertholletiae AJ287151 AF157243 AF113421 AF113459 C. echinulata AJ287152 AF157244 AF157130 AF157184 Dichotomocladium elegans AJ287153 AF157245 AF157131 AF157185 Gongronella butleri AJ287160 AF157252 AF157137 AF157191 Halteromyces radiatus AJ287161 AF157253 AF157138 AF157192 Hesseltinella vesiculosa AJ287163 AF157255 AF157140 AF157194 Lentamyces parricida AY944761 EU736250 EU736277 EU736304 L. zychae EU736228 EU736255 EU736282 EU736309 Lichtheimia corymbifera AJ287134 AF157227 AF113407 AF113445 L. hyalospora AJ287132 AF157225 AF157117 AF157171 L. hyalospora EF030531 EU826384 EU826360 EU826368 EU826377a EU826382a EU826361a EU826370a L. ramosaa Mortierella alpina EU736236 EU736263 EU736290 EU736317 M. multidivaricata AJ287168 AF157260 AF157144 AF157198 M. verticillata AJ287170 AF157262 AF157145 AF157199 Mucor racemosus AJ287177 AF157268 AF113430 AF113471 Mycotypha africana AJ287180 AF157271 AF157147 AF157201 M. microspora AJ287181 AF157272 AF157148 AF157202 Parasitella parasitica AJ287182 AF157273 AF157149 AF157203 Phycomyces blakesleeanus AJ287184 AF157275 AF157151 AF157205 Radiomyces spectabilis AJ287190 AF157281 AF157157 AF157211 Rhizomucor miehei AJ287191 AF157282 AF113432 AF113473 R. pusillus AJ287192 AF157283 AF113433 AF113474 Rhizopus oryzae AJ287198 AF157289 AF113440 AF113481 R. stolonifer AJ287199 AF157290 AF113441 AF113482 Saksenaea vasiformis AJ287200 AF157291 AF113442 AF113483 Spinellus fusiger AJ287201 AF157292 AF157159 AF157213 Thermomucor indicae-seudaticae AJ287208 AF157299 AF157165 AF157219 Umbelopsis isabellina AJ287209 AF157300 AF157166 AF157220 U. nana AJ287210 AF157301 AF157167 AF157221 U. ramanniana AJ287166 AF157258 X89435 AF113463 a described as Absidia idahoensis var. thermophila (Chen and Zheng 1998); sequences generated in this study 19 Identification of the Genus Absidia 457 Forty-two taxa were included, using three species of Mortierella as outgroup (Table 19.3). Starting from a random tree, two runs, each with four chains, were conducted for 5,000,000 generations. Thousand trees were sampled per run. The consensus tree was calculated using the halfcompat option with a 25% burn-in. The node confidence values (posterior probabilities, in percent) are shown above the branches in Fig. 19.1. 19.9.3 PCR-RFLP and Sequence Analysis ITS sequences were retrieved from GenBank (www.ncbi.nlm.nih.gov). Sequences for the genus Lichtheimia: L. blakesleeana (AY944892-4, EF030530), L. corymbifera/ramosa (DQ118984, DQ118982, AY944895, AY944897, AY944896); genus Lentamyces: L. parricida (AY944884/5), L. zychae (EF030529/AJ968561); genus Absidia sensu stricto: A. californica (AY944872/3), A. caerulea (AY944866-71), A. macrospora (AY944882/3), A. glauca (AY944875-81), A. spinosa (AY944886-8), A. repens (EF030527/8, AY944890/1, AJ877962, FJ849793), A. pseudocylindrospora (EF030525/6); A. anomala (EF030523), A. cylindrospora (AY944889), A. psychrophilia (AY944874), A. cuneospora (EF030524); genus Siepmannia: S. pineti (AJ748134), S. lariceti (AJ748857); Mucor circinelloides (AJ878535); Fennellomyces linderi (AJ878536). In silico restriction analysis as well as calculation of sequence similarities was done using BioEdit v.7.0.9.0 (Hall 1999). Acknowledgments This work was supported by a grant of the Deutsche Forschungsgemeinschaft (VO 772/9-1) and by the Thüringer Ministerium für Wissenschaft, Forschung und Kunst. The author wishes to express her gratitude to Kerstin Voigt (University of Jena, Germany) and Grit Walther (CBS, Centraalbureau voor Schimmelcultures Utrecht, The Netherlands) for critically reading the manuscript and valuable advices, to Keith A. Seifert (Eastern Cereal and Oilseed Research Centre, Agriculture and Agri-Food Ottawa, Ontario, Canada) for kindly providing a culture and sequence information of Absidia repens KAS3611, to Ru-yong Zheng (Key Laboratory of Systematic Mycology and Lichenology, Institute of Microbiology, Chinese Academy of Sciences, Beijing, China) for providing Absidia idahoensis var. thermophila strain AS 3.4808 and to Gisela Baumbach for excellent assistance in strain maintenance. References Abdel-Fattah AF, Ismail AMS, El-Aasar SA (1984) Production of rennin-like enzyme by Absidia cylindrospora. Agric Waste 11:125–131 Alastruey-Izquierdo A, Hoffmann K, Hoog de GS, Rodriguez-Tudela JL, Voigt K, Bibashi E, Walther G (2010) Species recognition and clinical relevance of the zygomycetous genus Lichtheimia (syn Mycocladus, Absidia spp.) in revision Bainier G (1882) Étude sur les Mucorinées. Thèse, Paris, France, p 136 Bainier G (1889) Sur l‘Absidia caerulea. Bull Soc Bot France 36:184 Bainier G (1903) Sur quelques espèces de Mucorinées nouvelles ou peu connues. Bull Soc Mycol France 19:153–172 Beauverie J (1900) Mycocladus verticillatus. Ann Univ Lyon 3:162–180 458 K. Hoffmann Brezezowska E, Dmochowska-Gladysz J, Kolek T (1996) Biotransformations XXXIX Metabolism of testosterone, androstenedione, progesterone and testosterone derivatives in Absidia coerulea culture. J Steroid Mol Biol 57:357–362 Chen GQ, Zheng RY (1998) A new thermophilic variety of Absidia idahoensis from China. Mycotaxon 69:173–179 Chen G, Yang M, Lu Z, Zhang J, Huang H, Liang Y, Guan S, Song Y, Wu L, Guo D (2007) Microbial Transformation of 20(S)-Protopanaxatriol-Type Saponins by Absidia coerulea. J Nat Prod 70:1203–1206 Cocconi G (1899) Ricerche intorno ad una nuova Mucorinea del genere Absidia van Tiegh. Accad Sci Bologna Mem Ser 5:85–90 Cohn FJ (1884) Mucor corymbifer. In: Lichtheim L (ed) Über pathogene Mucorineen und die durch sie erzeugten Mykosen des Kaninchens. Z Klin Med 7:149–177 Crous PW, Gams W, Stalpers JA, Robert V, Stegehuis G (2004) MycoBank: an online initiative to launch mycology into the 21st century. Stud Mycol 50:19–22 Dai T, Tegos GP, Burkatovskaya M, Castano AP, Hamblin MR (2009) Chitosan acetate bandage as a topical antimicrobial dressing of infected burns. Antimicrob Agents Chemother 53:393–400 Demirci F, Noma Y, Kirimer N, Can Başer KH (2004) Microbial Transformation of (-)-Carvone. Z Naturforsch 59:389–392 Ellis JJ, Hesseltine CW (1965) The genus Absidia: globose-spored species. Mycologia 57:222–235 Ellis JJ, Hesseltine CW (1966) Species of Absidia with ovoid sporangiospores II. Sabouraudia 5:57–77 Gonzalez CE, Rinaldi MG, Sugar AM (2002) Zygomycosis. Infect Dis Clin North Am 16:895–914 Guiraud P, Bonnet JL, Boumendjel A, Kadri-Dakir M, Dusser M, Bohatier J, Steiman R (2008) Involvement of Tetrahymena pyriformis and selected fungi in the elimination of anthracene, and toxicity assessment of the biotransformation products. Ecotoxicol Environ Saf 69:296–305 Hagem O (1908) Untersuchungen über norwegische Mucorineen. Math-Nat Kl 7:1–50 Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp 41:95–98 Hesseltine CW, Ellis JJ (1961) Notes on Mucorales, especially on Absidia. Mycologia 53:406–426 Hesseltine CW, Ellis JJ (1964) The genus Absidia: Gongronella and cylindrical-spored species of Absidia. Mycologia 56:568–601 Hesseltine CW, Ellis JJ (1966) Species of Absidia with ovoid sporangiospores I. Mycologia 58:761–785 Hibbett DS, Binder M, Bischoff JF, Blackwell M, Cannon PF, Eriksson OE, Huhndorf S, James T, Kirk PM, Lücking R, Lumbs HT, Lutzoni F, Matheny PB, McLaughlin DJ, Powell MJ, Redhead S, Schoch CL, Spatafora JW, Stalpers JA, Vilgalys R, Aime MC, Aptroot A, Bauer R, Begerow D, Benny GL, Castlebury LA, Crous PW, Dai YC, Gams W, Geiser DM, Griffith GW, Gueidan C, Hawksworth DL, Hestmark G, Hosaka K, Humber RA, Hyde KD, Ironside JE, Kõljalg U, Kurtzman CP, Larsson KH, Lichtwardt R, Longcore J, Miadlikowska J, Miller A, Moncalvo JM, Mozley-Standridge S, Oberwinkler F, Parmasto E, Reeb V, Rogers JD, Roux C, Ryvarden L, Sampaio JP, Schüssler A, Sugiyama J, Thorn RG, Tibell L, Untereiner WA, Walker C, Wang Z, Weir A, Weiss M, White MM, Winka K, Yao YJ, Zhang N (2007) A higher-level phylogenetic classification of the fungi. Mycol Res 111:509–547 Hoffmann K, Voigt K (2008) Absidia parricida plays a dominant role in biotrophic fusion parasitism among mucoralean fungi (Zygomycetes): Lentamyces, a new genus for A parricida and A zychae. Plant Biol 11:537–554 Hoffmann K, Discher S, Voigt K (2007) Revision of the genus Absidia (Mucorales, Zygomycetes): based on physiological, phylogenetic and morphological characters: thermotolerant Absidia spp. form a coherent group, the Mycocladiaceae fam nov. Mycol Res 111:1169–1183 Hoffmann K, Telle S, Walther G, Eckart M, Kirchmair M, Prillinger H, Prazenica A, Newcombe G, Dölz F, Papp T, Vágvölgyi C, deHoog S, Olsson L, Voigt K (2009a) Diversity, genotypic identification, ultrastructural and phylogenetic characterization of zygomycetes from different ecological habitats and climatic regions: limitations and utility of nuclear ribosomal DNA 19 Identification of the Genus Absidia 459 barcode markers. In: Gherbawy Y, Mach RL, Rai M (eds) Current advances in molecular mycology. Nova Science Publishers, Inc, New York, pp 263–312 Hoffmann K, Walther G, Voigt K (2009b) Mycocladus vs Lichtheimia: a correction (Lichtheimiaceae fam. nov., Mucorales, Mucoromycotina). Mycol Res 113:277–278 Huelsenbeck JP, Ronquist F (2001) MRBAYES: bayesian inference of phylogeny. Bioinformatics 17:754–755 Huszcza E, Dmochowska-Gladysz J (2003) Transformation of testosterone and related steroids in Absidia glauca culture. J Basic Microbiol 43:113–120 Kirk PM, Cannon PF, Minter DW, Stalpers JA (2008) Dictionary of the fungi, 10th edn. CABI, Wallingford, United Kingdom Kwaśna H, Nirenberg HI (2008a) Siepmannia, a new genus in the Mucoraceae. Mycologia 100:259–271 Kwaśna H, Nirenberg HI (2008b) Validation of the genus Siepmannia (Mucoraceae) and its four species. Polish Bot J 53:187 Kwaśna H, Ward E, Bateman GL (2006) Phylogenetic relationships among Zygomycetes from soil based on ITS1/2 rDNA sequences. Mycol Res 110:501–510 Lendner A (1907) Sur quelques Mucorinées. Bull Her Boiss, Sér II 7:249–251 Lendner A (1908) Les Mucorinées de la Suisse. Mat Flore cryptogamique Suisse 3:1–180 Lendner A (1924) Une Mucorinée nouvelle du genre Absidia. Bull Soc Bot Genève, Ser II 15:147–152 Machouart M, Larché J, Burton K, Collomb J, Maurer P, Cintrat A, Biava MF, Greciano S, Kuijpers AFA, Contet-Audonneau N, deHoog GS, Gérard A, Fortier B (2006) Genetic identification of the main opportunistic Mucorales by PCR-Restriction Fragment Length Polymorphism. J Clin Microbiol 44:805–810 McNeill J, Barrie FR, Burdet HM, Demoulin V, Hawksworth DL, Marhold K, Nicolson DH, Prado J, Silva PC, Skog JE, Wiersema JH, Turland NJ (eds) (2006) International code of botanical nomenclature (Vienna code). Regnum Vegetabile 146 ARG Ganter Verlag KG, Ruggell, Lichtenstein Mirza JH, Khan SM, Begum S, Shagufta S (1979) Mucorales of Pakistan. University of Agriculture, Faisalabad, Pakistan, p 183 Muzzarelli RAA, Illari P, Tarsi R, Dubini B, Xia W (1994) Chitosan from Absidia coerulea. Carbohydr Polym 25:45–50 O’Donnell K, Lutzoni FM, Ward TJ, Benny GL (2001) Evolutionary relationships among mucoralean fungi (Zygomycota): Evidence for family polyphyly on a large scale. Mycologia 93:286–296 Ribaldi M (1952) Sopra un interessante Zigomicete terricola: Gongronella urceolifera n gen., et n sp. Rivista di Biologia 44:157–166 Ribes JA, Vanover-Sams CL, Baker DJ (2000) Zygomycetes in human disease. Clin Microbiol Rev 13:236–301 Robert V, Stegehuis G, Stalpers J (2005) The MycoBank engine and related databases. http:// www.mycobank.org Ronquist F, Huelsenbeck JP (2003) MRBAYES 3: bayesian phylogenetic inference under mixed models. Bioinformatics 19:1572–1574 Rungsardthong V, Wongvuttanakul N, Kongpien N, Chotiwaranon P (2006) Application of fungal chitosan for clarification of apple jiuce. Process Biochem 41:589–593 Schipper MAA (1990) Notes on Mucorales – I. Observations on Absidia. Persoonia 14:133–149 Schwarz P, Bretagne S, Gantier JC, Garcia-Hermoso D, Lortholary O, Dromer F, Dannaoui E (2006) Molecular identification of Zygomycetes from culture and experimentally infected tissues. J Clin Microbiol 44:340–349 Sydow H (1903) Referate und kritische Besprechungen. Annals Mycologici 1:371 Thirion-Delalande C, Guillot J, Jensen HE, Crespeau FL, Bernex F (2005) Disseminated acute concomitant aspergillosis and mucormycosis in a pony. J Vet Med A Physiol Pathol Clin Med 52:121–124 460 K. Hoffmann Van Tieghem P (1876) Troisiéme mémoire sur les Mucorinées. Ann Sci Nat Bot 4:312–399 Vánová M (1991) Nomen novum, nomenclatural changes and taxonomic rearrangements in Mucorales. Ceská Mykologie 45:25–26 Voigt K, Wöstemeyer J (2001) Phylogeny and origin of 82 zygomycetes from all 54 genera of the Mucorales and Mortierellales based on combined analysis of actin and translation elongation factor EF-1alpha genes. Gene 270:113–120 Voigt K, Cigelnik E, O’Donnell K (1999) Phylogeny and PCR identification of clinical important Zygomycetes based on nuclear ribosomal-DNA sequence data. J Clin Microbiol 37:3957–3964 Voigt K, Hoffmann K, Einax E, Eckart M, Papp T, Vágvölgyi L, Olsson L (2009) Revision of the family structure of the Mucorales (Mucoromycotina, Zygomycetes) based on multigenegenealogies: phylogenetic analyses suggest a bigeneric Phycomycetaceae with Spinellus as sister group to Phycomyces. In: Gherbawy Y, Mach RL, Rai M (eds) Current advances in molecular mycology. Nova Science Publishers, New York, pp 263–312 Von Arx JA (1982) On Mucoraceae s. str. and other families of the Mucorales. Sydowia 35:10–26 Vuillemin P (1903) Le genre Tieghemella et la série de Absidées. Bull Soc Mycol France 19:119–127 Vuillemin P (1904) Le Lichtheimia ramosa (Mucor ramosus Lindt) champignon pathogène, distinct du L. corymbifera. Arch Parasitol 8:562–572 Zycha H (1935) Pilze II. Mucorineae. Krypt Fl Mark Brandenburg 6a, Leipzig Zycha H, Siepmann R, Linnemann G (1969) Mucorales. Eine Beschreibung aller Gattungen und Arten dieser Pilzgruppe, J Cramer, Lehre Chapter 20 Molecular Characters of Zygomycetous Fungi Xiao-yong Liu and Kerstin Voigt Abstract The traditional Zygomycota has recently been considered polyphyletic as evidenced by a lot of molecular phylogenetic analyses. As a result, it has been distributed into a new phylum and four pending subphyla. Before the taxonomic status for these four subphyla could be determined, the term “zygomycetous fungi” is used for those members traditionally included in the classical phylum Zygomycota. Most current molecular characters of zygomycetous fungi have revealed that there is an obvious conflict between the traditional morphology-based classification scheme and recent DNA-based phylogenies. Except for the notable adjustments at the phylum and subphylum level, major changes at the order level can be observed for Amoebidiales, Basidiobolales, Eccrinales, Entomophthorales, Geosiphonales and Mortierellales. With respect to families, studies on the order Mucorales have suggested an unnatural feature for its traditional family-level classification scheme. Some genera such as Absidia, Cunninghamella and Rhizopus have also been intensively investigated by molecular methods. Genes encoding glucoamylases, polygalacturonases, fumaric acids and polyunsaturated fatty acids, have been intensively studied for industrial purposes. Another important area is the study of the clinical relevance of zygomycetous fungi as pathogens. The poor sensitivity of histological practices, the difficult pure cultivation, and the inaccurate susceptibility and serological tests, have led to the development of highly sensitive and specific molecular techniques, such as microsatellite, oligonucleotide probes, microarrays of gene markers and their expression, fluorescent capillary electrophoresis, realtime PCR (polymerase chain reaction), PCR-RFLP (PCR-restriction fragment X-y. Liu Key Laboratory of Systematic Mycology and Lichenology, Institute of Microbiology, Chinese Academy of Sciences, No. 1 Beichen West Road, Chaoyang District, Beijing 100101, P. R. China e-mail: liuxiaoyong@im.ac.cn K. Voigt Institute of Microbiology, School of Biology and Pharmacy, University of Jena, Neugasse 25, 07743 Jena, Germany e-mail: kerstin.voigt@uni-jena.de Y. Gherbawy and K. Voigt (eds.), Molecular Identification of Fungi, DOI 10.1007/978-3-642-05042-8_20, # Springer-Verlag Berlin Heidelberg 2010 461 462 X-y. Liu and K. Voigt length polymorphism), RAPD (randomly amplified polymorphic DNA), PFGE (Pulsed Field Gel Electrophoresis), and direct sequencing of PCR products, but these methods are not widely available and are reserved primarily for research purposes. New techniques in the molecular identification of zygomycetous fungi need to be further developed and validated. So far, there are only five genome projects relevant to zygomycetous fungi, including Mortierella verticillata, Mucor circinelloides, Phycomyces blakesleeanus, Rhizopus arrhizus, and Smittium culisetae. More genome projects about industrially, agriculturally, medically and environmentally important zygomycetous fungi are hopeful to provide a better understanding of their natural status in the whole organismic system in the world and their potential to benefit the human being. 20.1 Introduction The phylum Zygomycota in the kingdom Fungi was first proposed by Moreau (1954), but it is invalid and consequently illegal because of a lack of Latin diagnosis or description, which is mandatory for descriptions on or after 1 January 1935 according to the International Code of Botanical Nomenclature (McNeill et al. 2006). Nevertheless, the name Zygomycota has still been recognized by most investigators for more than a semicentury and its members have been gradually increasing until the prosperity of molecular phylogenetic studies triggered by the proposal of universal primers for different rRNA genes and spacers in the early 1990s (White et al. 1990). Recently, the majority of molecular evidences strongly suggested that the traditional Zygomycota is not monophyletic (Tanabe et al. 2005; James et al. 2006), though its monophyletic feature was supported by the result of analyzing the sequences of genes rpb1 and rpb2 (Liu et al. 2006). In the most recent classification of the kingdom Fungi in light of numerous molecular phylogenetic studies, the widely accepted phylum Zygomycota was temporarily rejected and the taxa that have formerly been placed in it were distributed into five parts, including the phylum Glomeromycota and four subphyla of uncertain position (incertae sedis), namely Entomophthoromycotina, Kickxellomycotina, Mucoromycotina, and Zoopagomycotina (Hibbett et al. 2007). A three-protein phylogeny, shown in Fig. 20.1, demonstrates the phylogenetic positions of the former three subphyla with the exception of the Zoopagomycotina based on concatenation of translation elongation factor 1alpha, actin and beta tubulin amino acid sequences shown in Table 20.1. The notable taxonomic change of dispersing Zygomycota into different clades reflects to a great extent the current progress on molecular systematics of this basal fungal phylum, even though significant questions are left behind, such as the pending attribution of the four subphyla mentioned above in the kingdom of Fungi. As compared with the molecular polyphyletic trait of the classical phylum Zygomycota, its diversity in ecological distribution was also well known for a long period of time, ever since the very beginning of the establishment of this 20 Molecular Characters of Zygomycetous Fungi 463 Fig. 20.1 Concatenated phylogenetic analysis based on a total of 1,262 aligned amino acid characters comprising 500 characters of the translation elongation factor 1 alpha (TEF), 323 characters of actin (ACT) and 439 characters of beta-tubulin (BTUB) from sixty five taxa (Table 20.1). Single alignments were carried out using ClustalX version 1.83 (Higgins and Sharp 1988, 1989; Thompson et al. 1997). Neighbor-joining with distance measure mean character 464 X-y. Liu and K. Voigt phylum. Mucorales and Kickxellales are saprobes, the order Entomophthorales comprises obligate parasites of insects, Dimargaritales and Zoopagales are obligate parasites of microorganisms, Asellariales and Harpellales are endocommensals in crustacean guts, and the species of the order Endogonales are ectomycorrhizae. Taking into account the ecological diversity and the molecular polyphyly as well as the invalid naming for the phylum rank, the organisms that have long been included in the traditional Zygomycota refer to zygomycetous fungi in this chapter. That term is far from a new name but is widely accepted by lots of investigators. Here the range of zygomycetous fungi will exclude what is now classified as Glomeromycota, which is a phylum of proven acceptance in the kingdom of Fungi (Schübler et al. 2001). Furthermore, the Eccrinales and the Amoebidiales will be excluded from the zygomycetous fungi sensu stricto. Both orders were considered as members of Trichomycetes and now have moved on to a protistan class, the Mesomycetozoea (Cafaro 2005). An extra point worth noticing here targets the Microsporidia, for which zygomycete origin was suggested (Keeling et al. 2000; Keeling 2003) besides the fact that this group will not be included into the zygomycetous fungi in this chapter because of the many evidences revealing its protistan and nonfungal origin influencing its nomenclatural status (Forget et al. 2002; Gill and Fast 2006; Liu et al. 2006). The morphological taxonomy of zygomycetous fungi was extensively investigated during the last 50 years by some experts (Benjamin 1959, 1966, 1979; Hesseltine and Ellis 1973; O’Donnell 1979; Benny 1982; Lichtwardt 1986; Humber 1989; Benny et al. 2001) and intensively probed by more mycologists with a certain genus as research interests: Absidia (Hesseltine and Ellis 1964); Cunninghamella (Zheng and Chen 2001); Mortierella (Gams 1977); Mucor (Schipper 1973, 1975, 1976, 1978); Rhizopus (Zheng et al. 2007) etc. These studies circumscribe large genera if compared with genera within the derived fungi, the Asco- and Basidiomycota. For more details on numerous other genera and higher ranks in the zygomycetous fungi as well, it is worth to visit the open website www.zygomycetes.org by Gerald L. Benny which provides a great diversity and a wonderful overview about the zygomycetes. Looking back into the history, besides morphology, many other molecular and chemical traits such as sterols, fatty acids, nucleic acids, and proteins, were also applied for the identification and the classification of zygomycetous fungi. An investigation of sterol composition of zygomycetous fungi suggested that orders can be distinguished by different sterol chemotypes (Weete and Gandhi 1997). Excitingly, six new ergosterols from a marine zygomycetous fungus were found, providing an evidence for the importance to study on organisms in its endemic habitat (Wang et al. 2008). Regarding fatty acids, there are a few studies, which < Fig. 20.1 (continued) difference was conducted with PAUP*v4.0b10 (Swofford 1998); negative branch lengths were prohibited. Bootstrap supports (BS) (Felsenstein 1985; 50% majority rule) were obtained by 1,000 bootstrap replicates of a neighbor-joining search using mean character differences as distance measure as implemented in PAUP*v4.0b10. The tree was rooted to the elongation factor Tu and to the cytoskeletal proteins FtsZ (homologous to tubulin) and MreB (homologous to actin), the latter two first discovered in the thermophilic eubacterium Thermotoga maritima (van den Ent et al. 2001), facilitating deep-level phylogenies beyond the divergence of prokaryotes and eukaryotes. Names printed in red colour indicate taxa designated to the zygomycetes 20 Molecular Characters of Zygomycetous Fungi 465 Table 20.1 Protein sequences retrieved from the International Nucleotide Sequence Database Collaboration and genome projects (see footnotes) with corresponding accession numbers for the corresponding nucleotide sequences. The aligned amino acid sequences were subjected to distance based phylogenetic analyses shown in Fig. 20.1 Genus GenBank acc. nos. ACT TEF BTUB Acanthamoeba V00002 AY582829 AY582853 Acrasis – AF190771 AF276945 Allomyces Unpublished Unpublished AY131269 Anaeromyces Unpublished Unpublished Unpublished Aphanomyces – EF370041 EF370043 Arabidopsis U41998 X16430 M20405 Aspergillus M22869 AB007770 M17519 Basidiobolus – DQ282610 AF162060 e_gw1.2.331.1 estExt_40375 estExt_C_90068 Batrachochytriuma Bigelowiella EF455788 AY729489 EF455767 Bombyx X05185 D13338 X74951 Candida X16377 M29934 M19398 Capsaspora AY724689 DQ403163 – Catenaria Unpublished Unpublished AY944844 Chytriomyces AY582841 AY582823 AY944845 Coemansia – DQ282615 AY944833 Conidiobolus – DQ275337 AF162058 Coprinus AB034637 AY881026 AB000116 Corallochytrium AY582844 X55973 AY582850 Dictyostelium X03283 DQ282609 AF030823 Drosophila AB003910 X06869 M20419 Encephalitozoon AF031701 NC_003231 AF297876 Entamoeba M19871 M92073 AF247192 Entomophthora EF434860 ABB84538 AY944832 Gallus X00182 L00677 M11442 Giardia L29032 D14342 X06748 Glugea – D84253 AF162084 Glycine J01298 X56856 M21296 Harpochytrium – AF450113 AF162079 Histoplasma U17498 U14100 M28359 Homo M10277 X03558 X00734 Hydra XP_002154696 D79977 XM_002161824 Leishmania L16961 XM_001682206 AF345947 Lichtheimia AJ287134 AF157227 L47261 Ministeria AY582846 AY582825 AY582851 Monoblepharis – AF450112 AY944851 Monosiga AY026072 AY026073 AY026071 Mortierella AJ287170 AF157262 AF162071 estExt_C_40513 estExt_C_40175 gw1.3.598.1 Mucora Naegleria AF101729 DQ295229 X81050 Neoparamoeba EU089662 FJ807261 – Neurospora U78026 D45837 M13630 Nosema AF031702 AY452734 AY138803 Nuclearia AY582845 AY582827 AY582852 Orpinomyces Unpublished Unpublished Unpublished Oryza X16280 AF030517 X79367 (continued) 466 X-y. Liu and K. Voigt Table 20.1 (continued) Genus ACT Phycomycesa estExt_Genewise 1.C_340013 Physarum X07792 Phytophthora M59715 Plasmodiophora AM411664 Pythium X76725 Reticulomyxa AJ132374 RO3G_14002.3 Rhizopusb Saccharomyces L00026 Schizophyllum AF156157 Smittium AY582840 Sphaeroforma AJ780965 Stephanopogon EF455777 Suillus AF155931 Trichoderma X75421 Trichomonas U63124 Trypanosoma M20310 Umbelopsis AJ287166 Zea J01238 Thermotoga NP_228398 a Genome projects at http://genome.jgi-psf.org/ b Genome project at http://www.broadinstitute.org GenBank acc. nos. TEF e_gw1.2.294.1 AF016243 AJ249839 AM411655 DQ911417 EU810334 RO3G_15351.3 M10992 X94913 AY582822 DQ403164 FJ807246 AY883429 Z23012 – U10562 AF157258 D45408 M27479 BTUB estExt_20134 M20191 U22050 AM411665 AF218256 X96477 RO3G_06151.3 V01296 X63372 AY944829 – EF455757 AY112730 Z15054 L05468 K02836 AF162073 X52878 U65944 were reviewed by Kock and Botha (1998) and Frisvad et al. (2008). Because of the inconsistent quality throughout the zygomycetous fungi, no studies on fatty acids at higher taxonomic levels, from phylum to family, have been reported so far. But it is the very inconsistence that makes fatty acids so useful in lower rank’s taxonomy. Therefore at or below genus level applying comparatively few numbers of isolates as representatives, fatty acids were proven to be a potential marker for identification and classification as different genera, subgenera, or even species that can exhibit distinctive profiles (Amano et al. 1992; Blomquist et al. 1992; Weete et al. 1998; Weete and Gandhi 1999; Batrakov et al. 2002, 2004). Moving on to the nucleotide and protein aspects, there are far more papers than on the previously mentioned characters. Using “Zygomycota” as inquiry, more than 5,000 references for literature can nowadays be retrieved from the National Center for Biotechnology Information (www.ncbi.nlm.nih.gov) alone. This article will make an effort to give a sketchy review from such a considerably large number of publications on nucleic acids as well as proteins. 20.2 Molecular Characters for Classifying Zygomycetous Fungi It usually tends to be called molecular phylogeny or equivalents when DNA data are used to identify and classify organisms because they are believed to contain original, genetic and evolutionary information. Most current DNA data 20 Molecular Characters of Zygomycetous Fungi 467 Table 20.2 Comparison of traditional Zygomycota and current zygomycetous fungi Hawksworth et al. (1995) Kirk et al. (2001) Hibbett et al. (2007) ZYGOMYCOTA ZYGOMYCOTA GLOMEROMYCOTA Zygomycetes Zygomycetes Glomeromycetes 1. Geosiphonales 1. Glomales 2. Glomales 1. Glomerales SUBPHYLA INCERTAE SEDIS Entomophthoromycotina 2. Entomophthorales 3. Entomophthorales 1. Entomophthorales Basidiobolaceae 4. Basidiobolales 3. Endogonales Mortierellaceae 4. Mucorales 5. Endogonales 6. Mortierellales 7. Mucorales Mucoromycotina 1. Endogonales 2. Mortierellales 3. Mucorales 5. Zoopagales 8. Zoopagales Zoopagomycotina 1. Zoopagales 6. Dimargaritales 7. Kickxellales 9. Dimargaritales 10. Kickxellales Kickxellomycotina 1. Dimargaritales 2. Kickxellales Trichomycetes 1. Asellariales 2. Harpellales 3. Amoebidiales 4. Eccrinales Trichomycetes 1. Asellariales 2. Harpellales 3. Asellariales 4. Harpellales 3. Eccrinales from zygomycetous fungi have revealed that there is obviously a conflict between traditional morphology-based classification scheme and recent DNA-based phylogenetic one. Above the level of orders this inconsistence has led to a distribution of traditional Zygomycota among five clades, including Glomeromycota, Mucoromycotina, Kickxellomycotina, Entomophthoromycotina, and Zoopagomycotina (Hibbett et al. 2007). From the comparison of the current zygomycetous fungi with the last two editions of Ainsworth and Bisby’s Dictionary of the Fungi (Table 20.2), four major points of status changes at the order level can be observed. 20.2.1 Amoebidiales and Eccrinales Eccrinales and Amoebidiales, sharing ecological niche — the arthropod gut with other orders of the class Trichomycetes, were thought to be members of this class (Lichtwardt 1986). At the same time there were lots of opposites concerning its affinity with Trichomycetes based on the lack of a septal pore and associated plug, the presence of dictyosomes (Moss 1999), the lack of chitin in their cell wall (Trotter and Whisler 1965), and the distant relationships in rDNA phylogenies (Benny and O’Donnell 2000; Ustinova et al. 2000; Mendoza et al. 2002; Cafaro 2005). Therefore they are now excluded from the zygomycetous fungi. However, both of the remaining orders of the former Trichomycetes, the Asellariales and the 468 X-y. Liu and K. Voigt Harpellales still remain among the zygomycetous fungi. Zygospores were reported in species of the Asellariales (Valle and Cafaro 2008). The Harpellales is phylogenetically closely related to the Kickxellales (Benny and O’Donnell 2000) and classified now to the Kickxellomycotina (Hibbett et al. 2007). 20.2.2 Basidiobolales Basidiobolus was originally placed in the family Basidiobolaceae of the order Entomophthorales, then raised as a separate order Basidiobolales (Cavalier-Smith 1998) and later adopted (Kirk et al. 2001), and end up now as not being placed in any higher taxa. Why is that? Basidiobolus is currently the only genus of nonzoospore forming fungi known to have a nucleus-associated organelle that contain microtubules, suggesting a potential affinity between Basidiobolus and Chytridiomycota, although the structure of the microtubules is somewhat different (McKerracher and Heath 1985). In the phylogeny based on SSU rDNA, Basidiobolus did form a clade with many chytrids (James et al. 2000), whereas it was grouped with some Zygomycetes in phylogenies of alpha-tubulin and beta-tubulin and SSU rDNA (Nagahama et al. 1995; Jensen et al. 1998; Keeling et al. 2000; Keeling 2003) and nested in a clade consisting of major zygomycetous fungi in a stronger phylogeny of six genes (James et al. 2006). Based on ACT-BTUB-TEF protein phylogenies Basidiobolus appears to be paraphyletic to the Entomophthoromycotina, but with no statistical branch (BS) support (Fig. 20.1). Ultra structural and molecular characteristics congruously suggested that Basidiobolus might be a transitional organism between chytrids and zygomycetous fungi. 20.2.3 Geosiphonales Geosiphon was initially only given an unassigned zygomycete position as a genus (Hawksworth et al. 1995), then became a member of Geomycetes (Ascomycota) (Cavalier-Smith 1998), and later returned to Zygomycetes but with a certain position of a separate order Geosiphonales (Kirk et al. 2001; Schübler et al. 2001). In the current classification it is once more apart from the zygomycetous fungi and belongs to the Glomeromycota only reaching a level of family Geosiphonaceae (Hibbett et al. 2007; CABI_BioScience and Research 2008). The reason for these remarkable changes in position is that Geosiphon is associated with cyanobacteria, so different from its related taxa in Glomeromycota, which are all symbionts of arbuscular mycorrhiza. This time it will be expected to stay more stably because a series of papers about the Glomeromycota always revealed a monophyly for the genus Geosiphon and other members of Glomeromycota as a whole, even if a polyphyletic feature among them has also been discovered on the basis of data from ribosomal RNA and some protein genes (Schübler et al. 2001; Schübler 2002; Redecker and Raab 2006; Walker et al. 2007). 20 Molecular Characters of Zygomycetous Fungi 469 20.2.4 Mortierellales Mortierellales was established by Cavalier-Smith (1998) based on a traditional family Mortierellaceae in Mucorales, and this elevation in taxonomic status has been supported by different molecular data including act, EF-1 alpha, and SSUITS-LSU rRNA genes (Gehrig et al. 1996; Voigt and Wöstemeyer 2001; Lutzoni et al. 2004; Kwasna et al. 2006). While Hibbett et al. (2007) proposed the phylogenetic alliance of the Mortierellales with the Mucorales unified in the Mucoromycotina, Voigt et al. (2009) discovered a more derived phylogenetic relationship between the Mortierellales and the Endogonales, which deserves the erection of a new subphylum, the Mortierellomycotina ined. 20.2.5 Mucorales In regard to families, the most remarkable studies are about the classification within the order Mucorales. The artificial or unnatural feature of the traditional familylevel classification schemes for this order has been presented according to the analyses of SSU, LSU rDNA, EF-1 alpha, and actin gene sequences (O’Donnell et al. 2001; Voigt and Wöstemeyer 2001). Another gene, rpb1, was also proposed as an alternative marker for Mucorales phylogenetic studies (Tanabe et al. 2004). As a consequence of these data, G. L. Benny later integrated almost all families of Mucorales, except Umbelopsidaceae, into a single family Mucoraceae sensu lato (http://www.zygomycetes.org/), which was considered plausible (White et al. 2006). However, the recognition of families in ancient fungal lineages such as mucoralean and other terrestrial fungi is rather tedious. For a rough estimation of the origin and the radiation of mucoralean and allied fungi the following minimum ages for the divergence of major clades may be used: (1) 1,000 million years for the radiation of the major eukaryotic clades fungi, Metazoa and plants (Simon et al. 1993; Hightower and Meagher 1986), (2) the split of the Metazoa from the fungi 944–965 million years ago and (3) the divergence of Asco- and Basidiomycota 452–500 million years ago including (4) 440 million years for basidiomycete radiation and (5) 240 million years for radiation of the Pezizomycotina (for (2)–(5) see Berbee and Taylor 2001; Taylor and Berbee 2006). The standard procedure for molecular clocks is to plot averaged distances against time (Li and Graur 1999; Wang et al. 1999). Now exploiting the molecular clock-like and linear evolution of actin (Hightower and Meagher 1986; Berbee and Taylor 2001; Taylor and Berbee 2006) and other protein-coding genes commonly used for the assessment of evolutionary distances and phylogenetic trees (see Fig. 20.1 in this chapter; Fig. 2 in Voigt and Wöstemeyer 2001), the Mucorales may have already originated in the Late Cambrian, approximately 530 million years ago, and the origin of the Mortierellales dates to the Mid Devonian, 360 million years ago. Consequently, the origin of the Mortierellales coincides with the manifestation the arbuscular endomycorrhizal fungi (Glomeromycota) 353–462 million years ago (Simon et al. 1993; 470 X-y. Liu and K. Voigt Redecker et al 2000), whereas the Mucorales diverged as basal fungal lineage, possibly before the appearance of the first land plants, 360–480 million years ago (Simon et al. 1993; Hightower and Meagher 1986; Kenrick and Crane 1997; Berbee and Taylor 2001; Taylor and Berbee 2006). These data qualify the Mucorales as one of the earliest groups of recent land fungi on Earth, emerging before the availability of terrestrial plants as carbon source. Support of that estimation is gained by the putative localization of the radiation of the true fungi in the Early Paleozoic era, about 650–700 million years ago, which agrees well with previous estimates (Margulis 1981; Berbee and Taylor 2001; Taylor and Berbee 2006). It has to be emphasized that those rough calculations gained minimum ages, which fit well with fossil records (e.g., Redecker et al. 2000). Since fossil records define the time point of the manifestation of a certain taxon, the minimum ages may be underestimated. The suggestion of the ancient divergence between Glomeromycota from the Dikaryomycota (Asco- and Basidiomycota) 600 million years ago according to Redecker et al. (2000) strengthens that hypothesis. Therefore, it can be concluded that the Mucorales are more heterogenous and justify more than just two families. An attempt to revise the family structure of the Mucorales based on four-locus phylogenies is shown in Chaps. 11 (Fig. 11.4) and 19 (Fig. 19.1). Besides these significant shifts about subphyla, orders and families resulted from modern molecular phylogenetic studies; there are some more emphases on different genera of zygomycetous fungi. 20.2.5.1 Absidia Absidia spp. are filamentous fungi that are cosmopolitan and ubiquitous in nature as common environmental contaminants. They are found in plant debris and soil, as well as being isolated from foods and indoor air environment. They often cause food spoilage. It can transform steroids and produce rennin-like components, whereas some species are opportunistic human pathogens. Absidia is characterized by the branched and grouped sporangiophores carrying pyriform and relatively small sporangia, and arising on stolons from points between the rhizoids, but not opposite the rhizoids as in Rhizopus. Zygospores are formed on opposed, more or less equal suspensors adorned with several appendages. Absidia was divided into two parts: the subgenus Absidia in which the zygospores are surrounded by suspensor appendages and the subgenus Mycocladus in which suspensor appendages are not produced (Hesseltine and Ellis 1964). On investigating the phylogenetic relationships between 16 Absidia species based on act and ITS1-5.8S-ITS2 rDNA sequences, a trichotomy relevant to mesophilic, thermotolerant, and mycoparasitic groups was reconstructed, which is concordant with the morphology of the zygospores (Hoffmann et al. 2007). Furthermore based on the phylogenetic coherence of mesophilic and thermotolerant Absidia species, as well as other distinct characteristics in morphology, the two groups were separated into two distinct genera and placed in different family, Absidia (Absidiaceae) for the mesophilic species and Mycocladus (as “Mycocladiaceae,” but orthographically correct 20 Molecular Characters of Zygomycetous Fungi 471 Mycocladaceae) for the thermotolerant species A. corymbifera, A. blakesleeana and A. hyalospora (Hoffmann et al. 2007). But the type of species of Mycocladus, M. verticillatus Beauverie was discovered to be a coculture between a mesophylic and a mycoparasitic species of the former Absidia, and thus, not congeneric with the other species of Mycocladus. Therefore, a new family was established for the thermotolerant Absidia spp., the Lichtheimiaceae typified with Lichtheimia corymbifera (Cohn) Vuill. (Hoffmann et al. 2009b). For a more detailed review on the classification and the identification of Absidia see Chap. 19. 20.2.5.2 Actinomucor Actinomucor has been used in the fermentation of sufu (Chinese cheese). Only two species, viz. A. elegans and A. taiwanensis, are generally accepted in this small genus. The former is widespread and has already been found in many countries, while the latter has been reported from China only. According to ITS rDNA and EF-1 alpha sequence data, A. taiwanensis was reduced to varietal rank under A. elegans as A. elegans var. meitauzae because A. taiwanensis is the same fungus as Mucor meitauzae which was published in 1937, before A. taiwanensis (Zheng and Liu 2005). 20.2.5.3 Cunninghamella Cunninghamella is a filamentous fungus found in soil, plant material, animal material, cheese, and Brazil nuts. In addition to being a common contaminant, it is an opportunistic fungus causing infections in immunocompromised hosts. Cunninghamella can transform pantoprazole and amoxapine and can also produce gamma-linolenic acid, chitin and chitosan. Classification of Cunninghamella has been based principally on morphology of the sporangial and zygosporic states, maximum growth temperature, mating compatibility and zygospore formation. In addition, ITS rDNA sequences has been used as an important reference for species and variety delimitation, leading to the recognition of 12 species and three varieties within the genus (Liu et al. 2001). 20.2.5.4 Pilaira All members of Pilobolaceae are coprophilous and produce phototropic, almost unbranched sporangiophores, which arise directly from substrate and terminate in dark hemispheric columellate sporangia with persistent, cutinized walls. The genus Pilaira is characterized by the absence of trophocysts and subsporangial swellings, which are present in both Pilobolus and Utharomyces. Multigene phylogenetic analyses supported the presence of trophocysts and subsporangial swellings as synapomorphic characters by the separation of Pilaira from the core Pilobolaceae sensus stricto (Voigt et al. 2009). Nine of the ten species documented by Index 472 X-y. Liu and K. Voigt Fungorum of CABI Bioscience seem to be unique and have only been recorded once in literature, while P. anomala is ubiquitous in Europe and America and has been reported many times (Zheng and Liu 2009). Morphological studies have been conducted with all the available strains: six and 15 respectively from China and NRRL, resulting in recognition of five taxa including two new species and one new combination (Zheng and Liu 2009). Molecular phylogeny of 21 worldwide available strains of Pilaira, including two new species recently proposed from China, was reconstructed by using ITS rDNA and pyrG gene sequences (Liu et al. In press). The two loci displayed different phylogenetic histories. Besides some complete or partial concordances with morphology, several disagreements were found suggesting that this genus is dynamic in lineage splitting. 20.2.5.5 Rhizomucor The genus Rhizomucor is ubiquitous and commonly found in soil, compost heaps, decaying fruit and vegetables. It produces highly efficient enzymes for flax retting, milk clotting and lignocellulose degeneration. It is often associated with animal diseases. As for human being, most species of Rhizomucor are opportunistic agents causing zygomycosis. For successful treatments, it is critical to quickly and accurately identify the pathogen and then promptly and precisely apply antimycotics. Morphologically, Rhizomucor is distinguished from Mucor by the presence of stolons and rhizoids. Rhizomucor hitherto comprises six species and one variety belonging to two groups by virtue of the maximum growth temperature, i.e., the thermophilic and the mesophilic species. All these taxa can be well delimitated by distinct characteristics besides the maximum growth temperature, and the key to the species and varieties of Rhizomucor were provided by Zheng and Jiang (1995). It seems very easy to distinguish Rhizomucor taxa from one another according to numerous, stable, reliable and distinct morphological and physiochemical features. However, the practical determinations of Rhizomucor species based on morphological observations and sometimes physiochemical tests involve limitations, frequently resulting in inaccurate identifications (Lukács et al. 2004). Moreover recent molecular data have brought forth some new questions. The comparison of morphology-based and DNA-based identifications suggested that some prior reports concerning R. pusillus based on traditional methods might even represent other zygomycetous fungi, such as Rhizopus and Mucor (Iwen et al. 2005; Kontoyiannis et al. 2005). Different genes have shown different phylogenies for the members of Rhizomucor. Nuclear SSU rDNA sequence data have revealed that Rhizomucor were polyphyletic, with the thermophilic R. miehei and R. pusillus being a sister group of the clade of Absidia corymbifera, and the mesophilic R. variabilis being nested within the clade of seven species of Mucor (Voigt et al. 1999). The LSU rDNA sequences demonstrated another phylogenetic relationship between Rhizomucor spp. and other members of Mucorales, i.e., R. miehei and R. pusillus related next to Syncephalastrum racemosum rather than A. corymbifera, and R. variabilis formed a clade with two species of Mucor 20 Molecular Characters of Zygomycetous Fungi 473 instead of seven. Exhaustive ITS1-5.8S-ITS2 phylogenies and identity matrices revealed a close relationship between R. variabilis and M. hiemalis (Hoffmann et al. 2009a). The combined data of SSU, LSU rDNA and EF-1 alpha sequences showed that the species R. pusillus related more closely to Thermomucor indicae-seudaticae than to S. racemosum or A. corymbifera (O’Donnell et al. 2001). Beside these questions about R. miehei, R. pusillus and R. variabilis, the recognition of R. tauricus as a distinct species was doubted by some investigators based on isoenzyme patterns, ITS-RFLP, and RAPD (Vágvölgyi et al. 1999; Vastag et al. 2000). In order to solve the problems met in morphological and physiochemical identification, molecular relationships based on cox1, cox2, cox3, pyrG, and SSU+ITS1+5.8 S+ITS2 rDNA sequences, were analyzed with a result of dichotomy relevant to mesophilic and thermophilic groups (Liu and Zheng 2008). 20.2.5.6 Rhizopus Members of Rhizopus are important as agents of food fermentation, agricultural and food spoilage, human mucormycosis, and industrial and medical biotechnology. They may occur as saprobes on plant debris, soil, and dung, or as air contaminants. Rhizopus is characterized by apophysate sporangia, stolons on aerial mycelia, and rhizoids developed from stolons and opposite sporangiophores. Rhizopus was divided into two groups (R. microsporus Group and R. stolonifer Group) and one species (R. oryzae Went & Prins. Geerl.) according to anamorphic morphology and growth temperature (Schipper 1984; Schipper and Stalpers 1984). A recent monographic study of this genus based on morphology, maximum growth temperature, mating compatibility, and molecular systematics has been conducted in which a total of 17 taxa including ten species and seven varieties were recognized from a global collection (Zheng et al. 2007). Concerning molecular studies on this genus, some taxa have been involved for aiming at the resolution of high-level phylogeny (Voigt et al. 1999; O’Donnell et al. 2001; Schwarz et al. 2006). ITS, 18S and 28S rDNA sequences were used to investigate the molecular phylogeny of this genus (Abe et al. 2003, 2006; Hoffmann et al. 2009a). For a more detailed review on the molecular ITS-based identification of Rhizopus and allied genera see Chap. 11. ITS rDNA and pyrG were also applied to support most of the morphological treatments made recently, including the rejection of the level of group between genus and species (Liu et al. 2007; Zheng et al. 2007). On the basis of nuclear ribosomal DNA sequence and pyrG data, the distant relationship between the two varieties of R. stolonifer, i.e., var. lyococcus and var. stolonifer, was observed but treated differently: not reclassifying these two taxa (Abe et al. 2006), modifying the taxonomical scheme of Schipper et al. with a new combination R. lyococcus (Liou et al. 2007), or R. lyococcus being recognized as a synonym of R. reflexus (Liu et al. 2007; Zheng et al. 2007). As for those varieties within R. arrhizus and R. microsporus, IGS rDNA, especially the short tandem repeat motifs, was found to be very useful for variety delimitation (Liu et al. 2008). The fumaric–malic acid 474 X-y. Liu and K. Voigt producer group of R. arrhizae was raised to a species level as R. delemar, differing from the lactic acid producer group and confirmed by analyses of ITS rDNA, lactate dehydrogenase B, actin, translation elongation factor-1alpha and AFLP (Amplified Fragment Length Polymorphism) (Abe et al. 2007). Thus, it is obvious that molecular studies mostly paid attention to those issues that have long been questioned by traditional methods and actually addressed many, if not all. Therefore the molecular data were proven to be a very important part of the characters of zygomycetous fungi, not a separate and independent aspect. Among these molecular markers, genes and spacers of ribosomal RNA are most investigated, due to their universal primers, suitable fragment length and moderate variation speed. 20.3 Molecular Characters in Industrial Zygomycetous Fungi It is well-known that there are many products from zygomycetous fungi playing an important role in industry, such as glucoamylase, polygalacturonase, fumaric acid and polyunsaturated fatty acids (PUFAs). In addition to identification and classification, this article will also touch on topics of molecular characters in industrial fields. 20.3.1 Glucoamylase Glucoamylase from Rhizopus arrhizus has long been of considerable importance to the fermentation and food industries for saccharification of starch/amylopectin to alcohol. Many commercial glucoamylase enzyme preparations are derived from Rhizopus owing to their nearly complete conversion of starch to glucose. Glucoamylase has been isolated and characterized from a number of Rhizopus species (Mertens and Skory 2007a, b). 20.3.2 Polygalacturonase Flax has widely been used in textiles, high quality papers and composites. Retting is one of the greatest problems in flax fiber production. Water-retting and dew-retting were used in the past but are discarded nowadays due to different disadvantages such as pollution and weather dependency. Alternative ideas for retting have been developed, such as chemical retting using chelating agents and enzymatic retting using suitable enzyme mixtures. The latter technique results in cleaner fibers of higher and more consistent quality. The retting efficiency varied considerably between different organisms. Rhizopus oryzae (=R. arrhizus) produced extracellular enzymes that could independently degrade the middle lamella. This zygomycete 20 Molecular Characters of Zygomycetous Fungi 475 is an ideal model system for studying the mechanisms of enzymatic retting of flax. An extracellular polygalacturonase is probably the key component in the retting system of R. oryzae. It was purified and characterized. The purified enzyme has a molecular mass of 37,436 Da from mass spectrometric determination, an isoelectric point of 8.4, and has nonmethylated polygalacturonic acid as its preferred substrate. Peptide sequences indicate that the enzyme belongs to family 28, in similarity with other polygalacturonases. It contains, however an N-terminal sequence absent in other fungal pectinases, but present in an enzyme from the phytopathogenic bacterium Ralstonia solanacearum (Zhang et al. 2005). Besides R. oryzae, other wellknown zygomycetous fungus producing flax-retting enzymes is Rhizomucor pusillus (Henriksson et al. 1999). 20.3.3 Fumaric Acids Fumaric acid is a naturally occurring organic acid. Many microorganisms produce fumaric acid in small amounts, as it is a key intermediate in the citrate cycle. Currently, fumaric acid is produced chemically from maleic anhydride. However, as petroleum prices are rising rather quickly, maleic anhydride as a, petroleum derivative has increased in price as well. This situation has caused a renewed interest in the fumaric acid production by fermentation. Zygomycetous fungi, including Rhizopus, Mucor, Cunninghamella, and Circinella species, are wellknown for their organic acid-producing capability and have been used in fermentation processes for fumaric acid production (Roa Engel et al. 2008). Among these strains, R. arrhizus is the best-producing one. According to the analyses of, ITS rDNA, lactate dehydrogenase B, actin, translation elongation factor-1alpha, genomewide AFLP, and organic acid production as well, R. arrhizus var. delemar (=R. delemar) was thought to be the proper name for R. arrhizus fumaric–malic acid producers (Abe et al. 2007). To date there is no report about the sequencing, cloning and characterization of enzymes relevant to metabolic pathways of fumaric acid. 20.3.4 Polyunsaturated Fatty Acids Polyunsaturated fatty acids play important roles as structural components of membrane phospholipids and as precursors of the eicosanoids of signaling molecules. All mammals synthesize such eicosanoids, which are involved in inflammatory responses, reproductive function, immune responses and regulation of blood pressure. Arachidonic acid (AA; 20:4n 6), as a representative n 6 PUFA, is the most abundant 20-carbon PUFA in humans; and it not only exhibits various regulation effects and physiological activities but also plays important roles in infant nutrition. 476 X-y. Liu and K. Voigt Eicosapentaenoic acid (EPA; 20:5n 3), as a representative n 3 PUFA, is beneficial for cardiovascular diseases and decreases platelet aggregation and blood pressure. The distinct functions of the two families make the ratio in the diet of n 6 and n 3 PUFAs important for inflammatory responses and cardiovascular health. Mortierella alpina 1S-4 can produce EPA through the n 3 PUFA biosynthetic pathway and AA through the n 6 PUFA biosynthetic pathway. Therefore, this fungus is a good model for analyzing a fatty acid desaturation system from both fundamental and applied viewpoints. The genes encoding o3-desaturase, the D9-desaturases, D12-desaturases, D6-desaturases and D5-desaturases involved in 20-carbon PUFA biosynthesis have been cloned from M. alpina 1S-4 (Sakuradani et al. 1999a, b, c, 2005; Sakuradani and Shimizu 2003). 20.4 Molecular Characters in Medical Zygomycetous Fungi Another important area of studies on zygomycetous fungi is medicine. Zygomycosis is a rare and opportunistic infection caused by fungi belonging to zygomycetous fungi. This type of invasive infections, both superficial and angioinvasive, is major medical complications in immunocompromised patients. Zygomycosis is frequently lethal if it is not detected early and treated only with high doses of amphotericin B, which is currently the main effective therapy for zygomycosis fungi but limited by severe nephrotoxic side effects. Mortality rates may be as high as 80% in infected transplant recipients. The recent rise in AIDS, cancer, diabetes, leukemia, lymphoma, solid organ or bone marrow transplants, immunosuppressive therapy, and broad-spectrum antimicrobial drugs, has increased the number of immunosupprimised and immunocompromised patients. Although Aspergillus and Candida are generally two most commonly infected agents in such patients, zygomycosis has increased significantly over the past decade. Zygomycetous fungi are now listed by hospitals as microorganisms responsible for frequent emerging infections (Chayakulkeeree et al. 2006). Due to its acute and rapid development, the prompt and precise identification of a pathogen becomes very crucial for appropriate and efficient treatments to decrease its mortality. Zygomycosis can be subdivided into mucoromycosis and entomophthoromycosis which are caused by members of the order Mucorales and Entomophthorales, respectively. Pathogenic Mucorales comprises the following ten genera: Absidia, Apophysomyces, Cokeromyces, Cunninghamella, Mortierella, Mucor, Rhizomucr, Rhizopus, Saksenaea, and Syncephalastrum. Among these, Absidia, Mucor, Rhizomucor, and Rhizopus are four most common isolated pathogens. Medical Entomophthorales only includes Conidiobolus and Basidiobolus. Histopathological examination of the tissues typically shows characteristic broad, hyaline, ribbon-like, wide-angled branching, pauciseptate irregular fungal hyphae (mucoromycosis), or shows broad fungal hyphae with sparsely found septum surrounded by eosinophilic granular material (entomophthoromycosis). Their morphological characteristics are so reduced that it is impossible to distinguish them 20 Molecular Characters of Zygomycetous Fungi 477 only by histopathological examination. It is quite easy to differentiate them by sporangia and other reproductive features after pure cultivation. However, it has not been widely adopted to diagnose agents through detailed morphological characteristics due to notorious difficulties in axenic culture from clinical specimens because hyphal elements may be rare in tissue specimens and they can lose their viability during the tissue homogenization prior to culturing. In addition, Antifungal susceptibility and serological tests usually cannot get accurate and consistent endpoints and consequently are not available for routine use. The poor sensitivity of histological practices, the difficult pure cultivation, and the inaccurate susceptibility and serological tests, have led to the development of highly sensitive and specific molecular techniques, such as microsatellite, oligonucleotide probes, array, fluorescent capillary electrophoresis, real-time PCR, PCR-restriction fragment length polymorphism (PCR-RFLP), random amplified polymorphic DNA (RAPD), pulsed field gel electrophoresis (PFGE), and direct sequencing of PCR products. These methods targeted either a single gene or a whole genome. 20.4.1 SSU/LSU rDNA Numerous targets within the fungal genome have been evaluated, with much of the current work using areas within the ribosomal RNA gene (rDNA) complex, especially SSU and LSU rDNA. A molecular database for 42 isolates representing all clinically important zygomycetous fungi was constructed from the SSU and LSU rDNA. And 13 taxon-specific PCR primers were designed for those taxa most commonly encountered in infections, according to the aligned LSU rDNA sequences, which was suggested to have the potential to be used in the PCR assay for rapid and accurate identification of the etiological zygomycoses (Voigt et al. 1999). A case, in which proven invasive infection caused by Cunninghamella bertholletiae was confirmed by a pan-fungal PCR assay using conserved primers binding to SSU rDNA and a specific biotin-labeled probe, was reported. However, because of the relatively high conservation of SSU rDNA, the probe can detect DNA not only from C. bertholletiae but also from Absidia glauca, C. elegans and C. polymorpha (Rickerts et al. 2001). Other cases caused by C. bertholletiae, however, were also reported not by simple probes but through direct DNA sequencing of the PCR products from serial serum samples (Kobayashi et al. 2004; Makoto 2004). A set of oligonucleotide probes based on SSU rDNA sequencing for the detection of common airborne or pathogenic fungi at the genus and species levels, including some zygomycetous organisms, was developed (Wu et al. 2003). The MicroSeq D2 large-subunit ribosomal DNA sequencing kit was used to detect filamentous fungi, but did not obtain a high rate of accurate identification among the zygomycetous fungi, and consequently suggested an additional work to determine which gene or combination of genes is needed for complete separation of genera and species (Hall et al. 2004). Additionally, a review paper highlighted the 478 X-y. Liu and K. Voigt discordance between conventional phenotypic characterization and identification using this kit (Greenberg et al. 2004). 20.4.2 ITS rDNA The ITS rDNA has also been widely used as targets to detect and identify human fungal pathogens. It plays a critical role in the development of functional rRNA, with sequence variations among species showing promise as signature regions for molecular assays. A rapid identification of fungi was recommended by using the ITS2 rDNA region and an automated fluorescent capillary electrophoresis system, which was thought to be a promising tool for the rapid diagnosis of invasive fungal infections, including zygomycosis (Turenne et al. 1999). The whole ITS rDNA was used as the basis of multiplex PCR by which human pathogenic Rhizopus species was genetically identified and detected (Nagao et al. 2005) and as the basis of an oligonucleotide array, which was developed to identify species of clinically important filamentous or dimorphic fungi (Hsiao et al. 2005). The multiplex ITS-PCR method provided a rapid, simple, and reliable alternative to conventional methods to identify common clinical fungal isolates, based on the testing pathogenic fungi directly from cultures with 100% sensitivity and specificity (Luo and Mitchell 2002). PCR-RFLP method, another molecular biology tool concerning ITS rDNA, was developed to identify the main Mucorales belonging to the genera Absidia, Mucor, Rhizopus, and Rhizomucor involved in human pathology at genus and species level (Machouart et al. 2006). The analysis of ITS rDNA sequences was validated as a reliable technique for identification of zygomycetous fungi to the species level by using 54 strains belonging to 16 species, including the most common pathogenic Rhizopus, Absidia, Mucor, and Rhizomucor (Schwarz et al. 2006; Hoffmann et al. 2009a). Among the genus Cunninghamella, C. bertholletiae has long been considered the only agent in human diseases, but recently Cunninghamella echinulata has been identified as another agent based on the result of the ITS rDNA sequences of isolates of C. bertholletiae, which are highly homological and are distinct from those of C. echinulata (Liu et al. 2001; Lemmer et al. 2002). However, it is also mentioned that the number of organisms, which could be amplified directly from mycelial fragments is relatively low, only about 50% (Luo and Mitchell 2002). 20.4.3 Cytochrome b Gene Besides different regions and spacers of ribosomal RNA genes, other genes are also occasionally used to differentiate zygomycetous pathogens. For example, real-time PCR assay was developed by using probes binding to a 167-bp conserved region of the multicopy zygomycete cytochrome b gene, to detect species of the genera Absidia, Apophysomyces, Cunninghamella, Mucor, Rhizopus, and Saksenaea in 20 Molecular Characters of Zygomycetous Fungi 479 culture and tissue samples. Based on the high sensitivity and specificity from various materials, it is concluded that the real-time PCR assay was useful for the rapid and accurate detection of zygomycetous fungi (Hata et al. 2008). 20.4.4 Whole-Genome Fingerprinting and Genotyping The microsatellite DNA fingerprinting confirmed the proposal of the new pathogen of C. echinulata (Lemmer et al. 2002) and the interstrain polymorphism of Apophysomyces elegans was examined by using microsatellite primers with the results of two groups according to their patterns (Chakrabarti et al. 2003). RAPD (Randomly Amplified Polymorphic DNA) analysis is also able to provide reproducible markers for strain identification. RAPD analysis of Rhizomucor strains showed R. miehei to be genetically more homogeneous than the diverse R. pusillus. RAPD markers described in these works could be utilized in further studies to identify clinical and environmental isolates of R. miehei and R. pusillus and to check the accuracy of the original species identifications (Vastag et al. 2000). The intraspecies variability of Rhizopus stolonifer and R. oryzae (=R. arrhizus) species was also examined by the RAPD method (Vágvölgyi et al. 2004). Although only a few R. oryzae strains were involved in that study, the RAPD analysis appeared to support the unity of the species R. oryzae, which was established with the incorporation of about 30 strains originally described as independent species. PFGE (Pulsed Field Gel Electrophoresis) is also a versatile tool for molecular typing and to reveal the genetic variability at species and intraspecies levels. The electrophoretic karyotypes of A. glauca strains, have been revealed by rotating field gel electrophoresis and the sexually compatible strains of the mating type pair A. glauca showed considerable differences in their electrophoretic karyotype (Kayser and Wöstemeyer 1991); While those of Mucor circinelloides f. lusitanicus strains were generated by contour-clamped homogeneous electric field gel-electrophoresis and most showed polymorphisms with a different main karyotype pattern correlated with each mating type. (Dı́az-Mı́nguez et al. 1999). Further possiblities to elucidate fingerprints and genotypes are AFLP (Amplified Fragment Length Polymorphism), RFLP (Restriction Fragment Length Polymorphism), PCR-RFLP and microsatellite PCR (Vastag et al. 2000; Vágvölgyi et al 2004). 20.4.5 Carbon Assimilation Profiles In addition to the methods for investigating different genes and whole genome, carbon assimilation profiles are sometime adopted to determine agents of zygomycosis. Fifty seven strains belonging to 15 species and varieties of zygomycetous fungi, including the genera Rhizopus, Absidia, Mucor, and Rhizomucor, was tested for intraspecies and interspecies variability based on their carbon assimilation profiles. It was concluded that the clustering of isolates based on their carbon 480 X-y. Liu and K. Voigt assimilation profiles was in accordance with DNA-based phylogeny of zygomycetous fungi and the carbon assimilation profiles allowed precise and accurate identification of most zygomycetous fungi to the species level (Schwarz et al. 2007). Except the previously reviewed published studies, there are also many researches in which most infections are identified just as zygomycosis without any species determination (Ribes et al. 2000; Eucker et al. 2001). Nevertheless, the direct DNA sequencing of the PCR products obtained from pan-fungal primers remains the most reliable way to precisely identify zygomycetous fungi, even to a species level. But molecular techniques for detection of zygomycetous fungi by PCR or other methods are not widely available and are reserved primarily for research purposes. New techniques in the molecular identification of zygomycetous fungi need to be further developed and validated before they are used in clinical practice. 20.5 Genome Projects for Zygomycetous Fungi The genome projects relevant to zygomycetous fungi are far less than those for Ascomycota and Basidiomycota, with a total of five species, namely Mortierella verticillata, Mucor circinelloides, Phycomyces blakesleeanus, Rhizopus oryzae (=Rhizopus arrizus), and Smittium culisetae (Table 20.3). The R. oryzae genome is the first zygomycetous fungus to be sequenced and now has been assembled, annotated, mapped, and released to public. Its mitochondrial sequence was assembled separately from the genomic one (www.broad.mit.edu/annotation/genome/rhizopus_oryzae/MultiHome.html). Mitochondrial genome of another strain of R. oryzae has also been accomplished (www.ncbi.nlm.nih.gov/sites/entrez?db=genomeprj&cmd =Retrieve &dopt=Overview&list_uids=13352; Seif et al. 2005). The mitochondrial genomes of two other zygomycetous fungi are also determined (Seif et al. 2005), that is, M. verticillata for which the genome sequencing is in progress (www.broad. mit.edu/node/575), and S. culisetae. The M. circinelloides genome assembly was completed and is prerelease annotating (mucorgen.um.es/), while the genome of P. blakesleeanus was released (genome.jgi-psf.org/Phybl1/Phybl1.home.html). Table 20.3 Genome statistics for zygomycetous fungi (N/A, not available) Taxa Isolates Nuclei Mitochondrion Sizes (Mb) GC (%) Genes Sizes (kb) GC (%) Genes Mortierella NRRL 6337 N/A N/A N/A 58.745 27.9 54 verticillata Mucor CBS 277.49 36.05 N/A N/A N/A N/A N/A circinelloides Phycomyces N/A 55.9 N/A 14792 N/A N/A N/A blakesleeanus Rhizopus oryzae RA 99–880 46.09 35.6 17713 61.76 26.36 19 DAOM 148428 N/A N/A N/A 54.178 26.2 51 Smittium culisetae 18–3 N/A N/A N/A 58.654 18.5 61 20 Molecular Characters of Zygomycetous Fungi 481 These zygomycetous fungi which have been selected as material for genome programs are all of certain importance. For example, Mucor circinelloides is a model system for Agrobacterium tumefaciens-mediated transformation (Nyilasi et al. 2005); Phycomyces blakesleeanus is also a model system not only for sensory physiology, but also for the regulation of the biosynthesis of the pigment betacarotene in fungi as well; And Rhizopus oryzae (=R. arrhizus) is the most important and representative agent of mucormycosis. 20.6 Prospect Zygomycetous fungi are usually distinguished mainly on numerous morphological traits. When the circumscription of certain taxa was controversial, other characters were called on as an auxiliary measure, including molecular ones. The morphological characters are undoubtedly the core of taxonomy of zygomycetous fungi either the past, the present or the future. On the other hand, further works on molecular characters of the zygomycetous fungi will expectably continue to increase, especially for those fungi that play a crucial role in medicine, agriculture and industry. To resolve the pending phylogenetic relationships among zygomycetous fungi based on more loci and more comprehensive samplings, will continue to be an important research advance at higher ranks like phylum, class and family. Meanwhile, at relatively low levels such as genus and species, on the basis of evaluation of suitable molecular markers, thoroughly and rationally integrating morphological and molecular characters for the identification and classification of zygomycetous fungi and even all cellular organisms is definitely an unchangeable trend, as already taken on in the effort to establish a worldwide organism barcode system (International Barcode of Life at www.dnabarcoding.org, and Consortium for the Barcode of Life at barcoding.si.edu). Alternative genes, such as the single copy genes Mcm7 (MS456) and Tsr1 (MS277) useful for both phylogenetics and systematics (Aguileta et al. 2008, Schmitt et al. 2009) will circumvent designation problems triggered by paralogies of multicopied protein-coding genes or highly repetitive ribosomal DNA and revolutionize the molecular identification of fungi and the zygomycetes. It is most likely in the near future to establish a worldwide collaborative system for fungal identification serving all fungal research communities and individuals, on the basis of tremendous web resources, such as Assembling the Fungal Tree of Life (aftol.org), Barcode of Life Data Systems (www.barcodinglife. org), GenBank (www.ncbi.nlm.nih.gov), Global Biodiversity Information Faculty (www.gbif.org), Index Fungorum (www.indexfungorum.org), MycoBank (www. mycobank.org), etc. With the spurt in sequencing technology, more and more important zygomycetous fungi are hopeful of next candidates for genome projects, following the five completed representatives. The genome projects about some industrially, agriculturally, medically and environmentally important zygomycetous fungi are bound to provide a better understanding for their natural status in the whole organism system in the world and their potential to serve the human being. 482 X-y. Liu and K. Voigt References Abe A, Oda Y, Asano K, Sone T (2006) The molecular phylogeny of the genus Rhizopus based on rDNA sequences. Biosci Biotechnol Biochem 70:2387–2393 Abe A, Oda Y, Asano K, Sone T (2007) Rhizopus delemar is the proper name for Rhizopus oryzae fumaric-malic acid producers. Mycologia 99:714–722 Abe A, Sone T, Sujaya IN, Saito K, Oda Y, Asano K, Tomita F (2003) rDNA ITS sequence of Rhizopus oryzae: its application to classification and identification of lactic acid producers. Biosci Biotechnol Biochem 67:1725–1731 Aguileta G, Marthey S, Chiapello H, Lebrun MH, Rodolphe F, Fournier E, Gendrault-Jacquemard A, Giraud T (2008) Assessing the performance of single-copy genes for recovering robust phylogenies. Syst Biol 57:613–627 Amano N, Shinmen Y, Akimoto K, Kawashima H, Amachi T, SS YH (1992) Chemotaxonomic significance of fatty acid composition in the genus Mortierella (Zygomycetes, Mortierellaceae). Mycotaxon 44:257–265 Batrakov SG, Konova IV, Sheichenko VI, Esipov SE, Galanina LA, Istratova LN (2002) Unusual fatty acid composition of cerebrosides from the filamentous soil fungus Mortierella alpina. Chem Phys Lipids 117:45–51 Batrakov SG, Konova IV, Sheichenko VI, Esipov SE, Galanina LA, Istratova LN, Sergeeva YE (2004) Lipids of the zygomycete Absidia corymbifera F-965. Phytochemistry 65:1239–1246 Benjamin RK (1959) The merosporangiferous Mucorales. Aliso 4:321–433 Benjamin RK (1966) The merosporangium. Mycologia 58:1–42 Benjamin RK (1979) Zygomycetes and their spores. In: Kendrick B (ed) The whole fungus the sexual-asexual synthesis. National Museum of Natural Science, National Museums of Canada and the Kananaskis Foundation, Ottawa, Canada, pp 573–616 Benny GL (1982) Zygomycetes. In: Parker SP (ed) Synopsis and classification of living organisms, vol I. McGraw-Hill Book Company, New York, USA, pp 184–195 Benny GL, Humber RA, Morton JB (2001) The Zygomycota: Zygomycetes. In: McLaughlin DJ, McLaughlin EG, Lemke PA (eds) The Mycota. Systematics and Evolution, vol 7A. Springer, Berlin, pp 113–146 Benny GL, O’Donnell K (2000) Amoebidium parasiticum is a protozoan, not a Trichomycete. Mycologia 92:1133–1137 Berbee ML, Taylor JW (2001) Fungal molecular evolution: gene trees and geologic time. In: McLaughlin DJ, McLaughlin EG, Lemke PA (Eds.) The Mycota vol. VII Part B Systematics and Evolution (Esser K, Lemke PA, eds.). Springer, Berlin, pp 229-245 Blomquist G, Anderson B, Anderson K, Brondz I (1992) Analysis of fatty acids. A new method for characterization of moulds. J Microbiol Methods 16:59–68 CABI_BioScience, CBS, Research L (2008) CABI Databases. At http://www.indexfungorum.org/ Names/fundic.asp. Accessed 25 Dec. 2008 Cafaro MJ (2005) Eccrinales (Trichomycetes) are not fungi, but a clade of protists at the early divergence of animals and fungi. Mol Phylogenet Evol 35:21–34 Cavalier-Smith T (1998) A revised six-kingdom system of life. Biol Rev 73:203–266 Chakrabarti A, Ghosh A, Prasad GS, David JK, Gupta S, Das A, Sakhuja V, Panda NK, Singh SK, Das S, Chakrabarti T (2003) Apophysomyces elegans: an emerging zygomycete in India. J Clin Microbiol 41:783–788 Chayakulkeeree M, Ghannoum MA, Perfect JR (2006) Zygomycosis: the re-emerging fungal infection. Eur J Clin Microbiol Infect Dis 25:215–229 Dı́az-Mı́nguez JM, López-Matas MA, Eslava AP (1999) Complementary mating types of Mucor circinelloides show electrophoretic karyotype heterogeneity. Curr Genet 36:383–389 Eucker J, Sezer O, Graf B, Possinger K (2001) Mucormycoses. Mycoses 44:253–260 Felsenstein J (1985) Confidence limits of phylogenies: an approach using the bootstrap. Evolution 39:783–791 20 Molecular Characters of Zygomycetous Fungi 483 Forget L, Ustinova J, Wang Z, Huss VAR, Lang BF (2002) Hyaloraphidium curvatum: a linear mitochondrial genome, tRNA editing, and an evolutionary link to lower fungi. Mol Biol Evol 19:310–319 Frisvad JC, Andersen B, Thrane U (2008) The use of secondary metabolite profiling in chemotaxonomy of filamentous fungi. Mycol Res 112:231–240 Gams W (1977) A key to the species of Mortierella. Persoonia 9:381–391 Gehrig H, Schussler A, Kluge M (1996) Geosiphon pyriforme, a fungus forming endocytobiosis with Nostoc (cyanobacteria), is an ancestral member of the Glomales: evidence by SSU rRNA analysis. J Mol Evol 43:71–81 Gill EE, Fast NM (2006) Assessing the Microsporidia-fungi relationship: combined phylogenetic analysis of eight genes. Gene 375:103–109 Greenberg RN, Scott LJ, Vaughn HH, Ribes JA (2004) Zygomycosis (mucormycosis): emerging clinical importance and new treatments. Curr Opin Infect Dis 17:517–225 Hall L, Wohlfiel S, Roberts GD (2004) Experience with the MicroSeq D2 large-subunit ribosomal DNA sequencing kit for identification of filamentous fungi encountered in the clinical laboratory. J Clin Microbiol 42:622–626 Hata DJ, Buckwalter SP, Pritt BS, Roberts GD, Wengenack NL (2008) Real-time PCR method for detection of Zygomycetes. J Clin Microbiol 46:2353–2358 Hawksworth DL, Kirk PM, Sutton BC, Pegler DN (1995) Ainsworth and Bisby0 s dictionary of the Fungi, 8th edn. International Mycological Institute, CAB International, Wallingford, Oxon, UK Henriksson G, Akin DE, Slomczynski D, Eriksson KE (1999) Production of highly efficient enzymes for flax retting by Rhizomucor pusillus. J Biotechnol 68:115–123 Hesseltine CW, Ellis JJ (1964) The genus Absidia: Gongronella and cylindrical-spored species of Absidia. Mycologia 56:568–601 Hesseltine CW, Ellis JJ (1973) Mucorales. In: Ainsworth GC, Sparrow FK, Sussman AS (eds) The fungi, vol 4b. Academic Press, New York, USA, pp 187–217 Hibbett DS, Binder M, Bischoff JF, Blackwell M, Cannon PF, Eriksson OE, Huhndorf S, James T, Kirk PM, Lucking R, Thorsten LH, Lutzoni F, Matheny PB, McLaughlin DJ, Powell MJ, Redhead S, Schoch CL, Spatafora JW, Stalpers JA, Vilgalys R, Aime MC, Aptroot A, Bauer R, Begerow D, Benny GL, Castlebury LA, Crous PW, Dai YC, Gams W, Geiser DM, Griffith GW, Gueidan C, Hawksworth DL, Hestmark G, Hosaka K, Humber RA, Hyde KD, Ironside JE, Koljalg U, Kurtzman CP, Larsson KH, Lichtwardt R, Longcore J, Miadlikowska J, Miller A, Moncalvo JM, Mozley-Standridge S, Oberwinkler F, Parmasto E, Reeb V, Rogers JD, Roux C, Ryvarden L, Sampaio JP, Schussler A, Sugiyama J, Thorn RG, Tibell L, Untereiner WA, Walker C, Wang Z, Weir A, Weiss M, White MM, Winka K, Yao YJ, Zhang N (2007) A higher-level phylogenetic classification of the fungi. Mycol Res 111:509–547 Hightower RC, Meagher RB (1986) The molecular evolution of actin. Genetics 114:315–332 Higgins DG, Sharp PM (1988) CLUSTAL: a package for performing multiple sequence alignment on a microcomputer. Gene 73:237–244 Higgins DG, Sharp PM (1989) Fast and sensitive multiple sequence alignments on a microcomputer. Comput Appl Biosci 5:151–153 Hoffmann K, Discher S, Voigt K (2007) Revision of the genus Absidia (Mucorales, Zygomycetes) based on physiological, phylogenetic, and morphological characters; thermotolerant Absidia spp. form a coherent group, the Mycocladiaceae fam. nov. Mycol Res 111:1169–1183 Hoffmann K, Telle S, Walther G, Eckart M, Kirchmair M, Prillinger H, Prazenica A, Newcombe G, Dölz F, Papp T, Vágvölgyi C, deHoog S, Olsson L, Voigt K (2009a) Diversity, genotypic identification, ultrastructural and phylogenetic characterization of zygomycetes from different ecological habitats and climatic regions: limitations and utility of nuclear ribosomal DNA barcode markers. In: Gherbawy Y, Mach RL, Rai M (eds) Current advances in molecular mycology. Nova Science Publishers, Inc, New York, pp 263–312 484 X-y. Liu and K. Voigt Hoffmann K, Walther G, Voigt K (2009b) Mycocladus vs Lichtheimia: a correction (Lichtheimiaceae fam. nov., Mucorales, Mucoromycotina). Mycol Res 113:277–278 Hsiao CR, Huang L, Bouchara JP, Barton R, Li HC, Chang TC (2005) Identification of medically important molds by an oligonucleotide array. J Clin Microbiol 43:3760–3768 Humber RA (1989) Synopsis of a revised classification for the Entomophthorales (Zygomycotina). Mycotaxon 34:441–460 Iwen PC, Freifeld AG, Sigler L, Tarantolo SR (2005) Molecular identification of Rhizomucor pusillus as a cause of sinus-orbital zygomycosis in a patient with acute myelogenous leukemia. J Clin Microbiol 43:5819–5821 James TY, Kauff F, Schoch CL, Matheny PB, Hofstetter V, Cox CJ, Celio G, Gueidan C, Fraker E, Miadlikowska J, Lumbsch HT, Rauhut A, Reeb V, Arnold AE, Amtoft A, Stajich JE, Hosaka K, Sung GH, Johnson D, O’Rourke B, Crockett M, Binder M, Curtis JM, Slot JC, Wang Z, Wilson AW, Schuszler A, Longcore JE, O’Donnell K, Mozley-Standridge S, Porter D, Letcher PM, Powell MJ, Taylor JW, White MM, Griffith GW, Davies DR, Humber RA, Morton JB, Sugiyama J, Rossman AY, Rogers JD, Pfister DH, Hewitt D, Hansen K, Hambleton S, Shoemaker RA, Kohlmeyer J, Volkmann-Kohlmeyer B, Spotts RA, Serdani M, Crous PW, Hughes KW, Matsuura K, Langer E, Langer G, Untereiner WA, Lucking R, Budel B, Geiser DM, Aptroot A, Diederich P, Schmitt I, Schultz M, Yahr R, Hibbett DS, Lutzoni F, McLaughlin DJ, Spatafora JW, Vilgalys R (2006) Reconstructing the early evolution of Fungi using a six-gene phylogeny. Nature 443:818–822 James TY, Porter D, Leander CA, Vilgalys R, Longcore JE (2000) Molecular phylogenetics of the Chytridiomycota supports the utility of ultrastructural data in chytrid systematics. Can J Bot 78:336–350 Jensen AB, Gargas A, Eilenberg J, Rosendahl S (1998) Relationships of the insect-pathogenic order Entomophthorales (Zygomycota, Fungi) based on phylogenetic analyses of nuclear small subunit ribosomal DNA sequences (SSU rDNA). Fungal Genet Biol 24:325–334 Kayser T, Wöstemeyer J (1991) Electrophoretic karyotype of the zygomycete Absidia glauca: evidence for differences between mating types. Curr Genet 19:279–284 Keeling PJ (2003) Congruent evidence from alpha-tubulin and beta-tubulin gene phylogenies for a zygomycete origin of Microsporidia. Fungal Genet Biol 38:298–309 Keeling PJ, Luker MA, Palmer JD (2000) Evidence from beta-tubulin phylogeny that Microsporidia evolved from within the fungi. Mol Biol Evol 17:23–31 Kenrick P, Crane PR (1997) The origin and early evolution of plants on land. Nature 389:33–39 Kirk PM, Cannon PF, David JC, Stalpers JA (2001) Ainsworth & Bisby’s dictionary of the fungi, 9th edn. CABI Publishing, Wallingford Kobayashi M, Togitani K, Machida H, Uemura Y, Ohtsuki Y, Taguchi H (2004) Molecular polymerase chain reaction diagnosis of pulmonary mucormycosis caused by Cunninghamella bertholletiae. Respirology 9:397–401 Kock JLF, Botha A (1998) Fatty acids in fungal taxonomy. In: Frisvad JC, Bridge PD, Arora DK (eds) Chemical fungal taxonomy. Marcel Dekker, New York, USA, pp 219–246 Kontoyiannis DP, Lionakis MS, Lewis RE, Chamilos G, Healy M, Perego C, Safdar A, Kantarjian H, Champlin R, Walsh TJ, Raad II (2005) Zygomycosis in a tertiary-care cancer center in the era of Aspergillus-active antifungal therapy: a case-control obervational study of 27 recent cases. J Infect Dis 191:1350–1360 Kwasna H, Ward E, Bateman GL (2006) Phylogenetic relationships among Zygomycetes from soil based on ITS1/2 rDNA sequences. Mycol Res 110:501–510 Lemmer K, Losert H, Rickerts V, Just-NüblingG SA, Kekrmann M, Tintelnot K (2002) Identification of Cunninghamella species by molecular methods. Mycoses 45:31–36 Li WH, Graur D (1999) Fundamentals of molecular evolution. Sinauer, Sunderland, MA Lichtwardt RW (1986) The Trichomycetes. Fungal associates of Arthropods. Springer, New York, USA, p 343 Liou GF, Chen SR, Wei YH, Lee FI, Fu HM, Yuan GF, Stalpers JA (2007) Polyphasic approach to the taxonomy of the Rhizopus stolonifer group. Mycol Res 111:196–203 20 Molecular Characters of Zygomycetous Fungi 485 Liu XY, Huang H, Zheng RY (2001) Relationships within Cunninghamella based on sequence analysis of ITS rDNA. Mycotaxon 80:77–95 Liu XY, Huang H, Zheng RY (2007) Molecular phylogenetic relationships within Rhizopus based on combined analyses of ITS rDNA and pyrG gene sequences. Sydowia 59:235–253 Liu XY, Huang H, Zheng RY (2008) Delimitation of Rhizopus varieties based on IGS rDNA sequences. Sydowia 60:93–112 Liu XY, Zheng RY (2008) Multiple loci resolving species of Rhizomucor and drawing attention to related Mucor 2008 International Symposium on Fungal Diversity. Hangzhou, China, Oct. 16–19, 2008. p 22 Liu YJ, Hodson MC, Hall BD (2006) Loss of the flagellum happened only once in the fungal lineage: phylogenetic structure of kingdom Fungi inferred from RNA polymerase II subunit genes. BMC Evol Biol 6:74. doi:10.1186/1471-2148-6-74 Lukács B, Papp T, Nyilasi I, Nagy E, Vágvölgyi C (2004) Differentiation of Rhizomucor species on the basis of their different sensitivities to lovastatin. J Clin Microbiol 42:5400–5402 Luo G, Mitchell TG (2002) Rapid identification of pathogenic fungi directly from cultures by using multiplex PCR. J Clin Microbiol 40:2860–2865 Lutzoni F, Kauff F, Cox CJ, McLaughlin D, Celio G, Dentinger B, Padamsee M, Hibbett D, James TY, Baloch E, Grube M, Reeb V, Hofstetter V, Schoch C, Arnold AE, Miadlikowska J, Spatafora J, Johnson D, Hambleton S, Crockett M, Shoemaker R, Sung GH, Lucking R, Lumbsch T, O’Donnell K, Binder M, Diederich P, Ertz D, Gueidan C, Hansen K, Harris RC, Hosaka K, Lim YW, Matheny B, Nishida H, Pfister D, Rogers J, Rossman A, Schmitt I, Sipman H, Stone J, Sugiyama J, Yahr R, Vilgalys R (2004) Assembling the fungal tree of life: progress, classification, and evolution of subcellular traits. Am J Bot 91:1446–1480 Machouart M, Larché J, Burton K, Collomb J, Maurer P, Cintrat A, Biava MF, Greciano S, Kuijpers AF, Contet-Audonneau N, de Hoog GS, Gérard A, Fortier B (2006) Genetic identification of the main opportunistic Mucorales by PCR-restriction fragment length polymorphism. J Clin Microbiol 44:805–810 Makoto KO (2004) Molecular polymerase chain reaction diagnosis of pulmonary mucormycosis caused by Cunninghamella bertholletiae. Respirology 9:397–401 Margulis L (1981) Symbiosis in cell evolution – Life and ist environment on the early earth. W.H. Freeman and Company, San Francisco McKerracher LJ, Heath IB (1985) The structure and cycle of the nucleus-associated organelle in two species of Basidiobolus. Mycologia 77:412–417 McNeill J, Barrie FR, Burdet HM, Demoulin V, Hawksworth DJ, Marhold K, Nicolson DH, Prado J, Silva PC, Skog JE, Wiersema JH, Turland NJ (2006) International code of botanical nomenclature (Vienna code) adopted by the seventh international botanical congress Vienna, Austria, July 2005. A. R. G, Gantner Verlag, Ruggell, Liechtenstein, p 568 Mendoza L, Taylor JW, Ajello L (2002) The class Mesomycetozoea: a heterogeneous group of microorganisms at the animal-fungal boundary. Annu Rev Microbiol 56:315–344 Mertens JA, Skory CD (2007a) Isolation and characterization of a second glucoamylase gene without a starch binding domain from Rhizopus oryzae. Enzyme Microb Tech 40:874–880 Mertens JA, Skory CD (2007b) Isolation and characterization of two genes that encode active glucoamylase without a starch binding domain from Rhizopus oryzae. Curr Microbiol 54:462–466 Moreau F (1954) Les Champignons Physiologie, morphologie, développment et systématique. Lechevalier, Paris Moss ST (1999) Astreptonema gammari: an eccrinid with appendaged spores. Kew Bull 54:637–650 Nagahama T, Sato H, Shimazu M, Sugiyama J (1995) Phylogenetic divergence of the entomophthoralean fungi: evidence from nuclear 18S ribosomal RNA gene sequences. Mycologia 87:203–209 Nagao K, Ota T, Tanikawa A, Takae Y, Mori T, Udagawa S, Nishikawa T (2005) Genetic identification and detection of human pathogenic Rhizopus species, a major mucormycosis 486 X-y. Liu and K. Voigt agent, by multiplex PCR based on internal transcribed spacer region of rRNA gene. J Dermatol Sci 39:23–31 Nyilasi I, Ács K, Papp T, Nagy E, Vágvölgyi C (2005) Agrobacterium tumefaciens-mediated transformation of Mucor circinelloides. Folia Microbiol 50:415–420 O’Donnell K (1979) Zygomycetes in culture. Palfrey Contributions in Botany. No 2. Department of Botany. University of Georgia, Athens, Georgia, USA, p 257 O’Donnell K, Lutzoni F, Ward TJ, Benny GL (2001) Evolutionary relationships among mucoralean fungi (Zygomycota): evidence for family polyphyly on a large scale. Mycologia 93:286–296 Redecker D, Kodner R, Graham LE (2000) Glomalean fungi from the Ordovician. Science 289:1920–1921 Redecker D, Raab P (2006) Phylogeny of the Glomeromycota (arbuscular mycorrhizal fungi): recent developments and new gene markers. Mycologia 98:885–895 Ribes JA, Vanover-Sams CL, Baker DJ (2000) Zygomycetes in human disease. Clin Microbiol Rev 13:236–301 Rickerts V, Loeffler J, Böhme A, Einsele H, Just-Nübling G (2001) Diagnosis of disseminated zygomycosis using a polymerase chain reaction assay. Eur J Clin Microbiol Infect Dis 20:744–745 Roa Engel CA, Straathof AJ, Zijlmans TW, van Gulik WM, van der Wielen LA (2008) Fumaric acid production by fermentation. Appl Microbiol Biotechnol 78:379–389 Sakuradani E, Abe T, Iguchi K, Shimizu S (2005) A novel fungal 3-desaturase with wide substrate specificity from arachidonic acid-producing Mortierella alpina 1 S–4. Appl Microbiol Biotechnol 66:648–654 Sakuradani E, Kobayashi M, Shimizu S (1999a) Delta 6-fatty acid desaturase from an arachidonic acid-producing Mortierella fungus: gene cloning and its heterologous expression in a fungus, Aspergillus. Gene 238:445–453 Sakuradani E, Kobayashi M, Shimizu S (1999b) Delta 9-fatty acid desaturase from arachidonic acid-producing fungus. Unique gene sequence and its heterologous expression in a fungus, Aspergillus. Eur J Biochem 260:208–216 Sakuradani E, Kobayashi M, Ashikari T, Shimizu S (1999c) Identification of Delta 12-fatty acid desaturase from arachidonic acid-producing Mortierella fungus by heterologous expression in the yeast Saccharomyces cerevisiae and the fungus Aspergillus oryzae. Eur J Biochem 261:812–820 Sakuradani E, Shimizu S (2003) Gene cloning and functional analysis of a second delta 6-fatty acid desaturase from an arachidonic acid-producing Mortierella fungus. Biosci Biotechnol Biochem 4:704–711 Schipper MAA (1973) A study on variability in Mucor hiemalis and related species. Stud Mycol 4:1–40 Schipper MAA (1975) Mucor mucedo, Mucor flavus and related species. Stud Mycol 10:1–33 Schipper MAA (1976) On Mucor circinelloides, Mucor racemosus and related species. Stud Mycol 12:1–40 Schipper MAA (1978) On certain species of Mucor with a key to all accepted species. Stud Mycol 17:1–52 Schipper MAA (1984) A revision of the genus Rhizopus I: the Rhizopus stolonifer-group and Rhizopus oryzae. Stud Mycol 25:1–19 Schipper MAA, Stalpers JA (1984) A revision of the genus Rhizopus II: the Rhizopus microsporusgroup. Stud Mycol 25:20–34 Schmitt I, Crespo A, Divakar PK, Fankhauser JD, Herman-Sackett E, Kalb K, Nelsen MP, Nelson NA, Rivas-Plata E, Shimp AD, Widhelm T, Lumbsch HT (2009) New primers for promising single copy genes in fungal phylogenetics and systematics. Persoonia 23:35–40 Schübler A (2002) Molecular phylogeny, taxonomy, and evolution of Geosiphon pyriformis and arbuscular mycorrhizal fungi. Plant Soil 244:75–83 Schübler A, Schwarzott D, Walker C (2001) A new fungal phylum, the Glomeromycota: phylogeny and evolution. Mycol Res 105:1413–1421 20 Molecular Characters of Zygomycetous Fungi 487 Schwarz P, Bretagne S, Gantier JC, Garcia-Hermoso D, Lortholary O, Dromer F, Dannaoui E (2006) Molecular identification of zygomycetes from culture and experimentally infected tissues. J Clin Microbiol 44:340–349 Schwarz P, Lortholary O, Dromer F, Dannaoui E (2007) Carbon assimilation profiles as a tool for identification of zygomycetes. J Clin Microbiol 45:1433–1439 Seif E, Leigh J, Liu Y, Roewer I, Forget L, Lang BF (2005) Comparative mitochondrial genomics in zygomycetes: bacteria-like RNase P RNAs, mobile elements and a close source of the group I intron invasion in angiosperms. Nucleic Acids Res 33:734–744 Simon L, Bousquet J, Lévesque RC, Lalonde M (1993) Origin and diversification of endomycorrhizal fungi and coincidence with vascular land plants. Nature 363:67–69 Swofford DL (1998) PAUP*: phylogenetic analysis using Parsimony (*and Other Methods) Version 4. Sinauer Associates, Sunderland, MA Tanabe Y, Saikawa M, Watanabe MM, Sugiyama J (2004) Molecular phylogeny of Zygomycota based on EF-1alpha and RPB1 sequences: limitations and utility of alternative markers to rDNA. Mol Phylogenet Evol 30:438–449 Tanabe Y, Watanabe MM, Sugiyama J (2005) Evolutionary relationships among basal fungi (Chytridiomycota and Zygomycota): insights from molecular phylogenetics. J Gen Appl Microbiol 51:267–276 Taylor JW, Berbee ML (2006) Dating divergences in the fungal tree of life: review and new analyses. Mycologia 98:838–849 Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG (1997) The Clustal X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res 25:4876–4882 Trotter MJ, Whisler HC (1965) Chemical composition of the cell wall of Amoebidium parasiticum. Can J Bot 43:869–876 Turenne CY, Sanche SE, Hoban DJ, Karlowsky JA, Kabani AM (1999) Rapid identification of Fungi by using the ITS2 genetic region and an automated fluorescent capillary electrophoresis system. J Clin Microbiol 37:1846–1851 Ustinova I, Krienitz L, Huss VAR (2000) Hyaloraphidium curvatum is not a green alga, but a lower fungus; Amoebidium parasiticum is not a fungus, but a member of the DRIPS. Protist 151:253–262 Valle LG, Cafaro MJ (2008) First report of zygospores in Asellariales and new species from the Caribbean. Mycologia 100:122–131 van den Ent F, Amos LA, Löwe J (2001) Prokaryotic origin of the actin cytoskeleton. Nature 413:39–44 Vastag M, Papp T, Kasza Z, Vágvölgyi C (2000) Intraspecific variation in two species of Rhizomucor assessed by random amplified polymorphic DNA analysis. J Basic Microbiol 40:269–277 Vágvölgyi C, Heinrich H, Ács K, Papp T (2004) Genetic variability in the species Rhizopus stolonifer, assessed by random amplified polymorphic DNA analysis. Antonie Leeuwenhoek 86:181–188 Vágvölgyi C, Vastag M, Ács K, Papp T (1999) Rhizomucor tauricus: a questionable species of the genus. Mycol Res 103:1318–1322 Voigt K, Cigelnik E, O’Donnell K (1999) Phylogeny and PCR identification of clinically important Zygomycetes based on nuclear ribosomal-DNA sequence data. J Clin Microbiol 37:3957–3964 Voigt K, Hoffmann K, Einax E, Eckart M, Papp T, Vágvölgyi L, Olsson L (2009) Revision of the family structure of the Mucorales (Mucoromycotina, Zygomycetes) based on multigenegenealogies: phylogenetic analyses suggest a bigeneric Phycomycetaceae with Spinellus as sister group to Phycomyces. In: Gherbawy Y, Mach RL, Rai M (eds) Current advances in molecular mycology. Nova Science Publishers, Inc, New York, pp 263–312 Voigt K, Wöstemeyer J (2001) Phylogeny and origin of 82 zygomycetes from all 54 genera of the Mucorales and Mortierellales based on combined analysis of actin and translation elongation factor EF-1alpha genes. Gene 270:113–120 488 X-y. Liu and K. Voigt Walker C, Vestberg M, Demircik F, Stockinger H, Saito M, Sawaki H, Nishmura I, Schüssler A (2007) Molecular phylogeny and new taxa in the Archaeosporales (Glomeromycota): Ambispora fennica gen. sp. nov., Ambisporaceae fam. nov., and emendation of Archaeospora and Archaeosporaceae. Mycol Res 111:137–153 Wang F, Fang Y, Zhang M, Lin A, Zhu T, Gu Q, Zhu W (2008) Six new ergosterols from the marine-derived fungus Rhizopus sp. Steroids 73:19–26 Wang DYC, Kumar S, Hedges SB (1999) Divergence time estimates for the early history of animal phyla and the origin of plants, animals and fungi. Proc Roy Soc Lond B 266:163–171 Weete JD, Gandhi S (1997) Sterols of the phylum Zygomycota: phylogenetic implications. Lipids 32:1309–1316 Weete JD, Gandhi SR (1999) Sterols and fatty acids of the Mortierellaceae: taxonomic implications. Mycologia 91:642–649 Weete J, Shewmaker F, Gandhi S (1998) g-Linolenic acid in zygomycetous fungi: Syzygites megalocarpus. J Am Oil Chem Soc 75:1367–1372 White MM, James TY, O’Donnell K, Cafaro MJ, Tanabe Y, Sugiyama J (2006) Phylogeny of the Zygomycota based on nuclear ribosomal sequence data. Mycologia 98:872–884 White TJ, Bruns T, Lee S, Taylor JW (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ (eds) PCR protocols: a guide to methods and applications. Academic Press, New York, USA, pp 315–322 Wu Z, Tsumura Y, Blomquist G, Wang XR (2003) 18S rRNA gene variation among common airborne fungi, and development of specific oligonucleotide probes for the detection of fungal isolates. Appl Environ Microbiol 69:5389–5397 Zhang J, Henriksson H, Szabo IJ, Henriksson G, Johansson G (2005) The active component in the flax-retting system of the zygomycete Rhizopus oryzae sb is a family 28 polygalacturonase. J Ind Microbiol Biotechnol 32:431–438 Zheng RY, Chen GQ (2001) A monograph of Cunninghamella. Mycotaxon 80:1–75 Zheng RY, Chen GQ, Huang H, Liu XY (2007) A monograph of Rhizopus. Sydowia 59:273–372 Zheng RY, Jiang H (1995) Rhizomucor endophyticus sp. nov., an endophytic zygomycetes from higher plants. Mycotaxon 56:455–466 Zheng RY, Liu XY (2005) Actinomucor elegans var. meitauzae, the correct name for A. taiwanensis and Mucor meitauzae (Mucorales, Zygomycota). Nova Hedwigia 80:419–431 Zheng RY, Liu XY (2009) Taxa of Pilaira (Mucorales, Zygomycota) from China. Nova Hedwigia 88:255–267 Index A Absidia, 464, 470–474, 478, 479 A. aegyptiacum, 452 A. anomala, 444, 451, 457 A. caerulea, 444, 451, 452, 457 A. californica, 444, 451, 457 A. capillata, 452 A. clavata, 452 A. corymbifera, 445 A. cuneospora, 444, 451, 457 A. cylindrospora, 444, 451, 452, 457 A. cylindrospora var. cylindrospora, 441, 451 A. cylindrospora var. nigra, 441, 451 A. cylindrospora var. rhizomorpha, 441, 451 A. dubia, 446 A. fassatiae, 452 A. fusca, 444, 451 A. glauca, 444, 451, 457 A. griseola, 452 A. heterospora, 451 A. idahoensis var. thermophila, 440 A. inflata, 452 A. macrospora, 444, 451, 457 A. narayanai, 452 A. pseudocylindrospora, 444, 451, 452 A. psychrophilia, 444, 452, 457 A. reflexa, 451, 452 A. repens, 444, 451, 454, 455, 457 A. robusta, 452 A. scabra, 451, 452 A. septata, 451, 452 A. spinosa, 444, 457 A. spinosa var. azygospora, 441, 451 A. spinosa var. biappendiculata, 441, 451 A. spinosa var. spinosa, 441, 451 A. tuneta, 452 A. ushtrina, 452 description, 440 distinction, 444, 445 mesophilic species, 440, 444 molecular aspects, 440 molecular key, 452–454 morphological aspects, 440–442 phylogenetic analyses, 442, 444, 453 phylogeny, 442 physiological aspects, 440, 442, 445 polyphyly, 442 sensu lato, 440–442 sensu stricto, 440–442, 444–445, 453, 457 synoptic key, 450–452 thermotolerant species, 440, 445, 449 Absidiaceae, 444–445, 450–452 Abundance, underestimated, 71 Acid arachidonic acid, 475 eicosapentaenoic acid (EPA), 476 fatty acids, 464, 466, 474–476 fumaric acids, 475 organic acid, 475 Actin, MreB, 463 Actinomucor, 471 489 490 Albuginales, 39 Albuginomycetidae, 37 Albugo candida, 37, 41 All Fungi Barcoding, 146 Alternative host plants, 73 Amplified fragment length polymorphism (AFLP), 40, 113, 134, 137, 198–200, 202, 204, 279–281 Anamorphs, 287 Ancient DNA, 35 Anisogramma anomala, 16 Antimycotics amphotericin B, 476 nephrotoxic side effects, 476 Antrodia spp., 253 Antrodia vaillantii, 262 Arabidopsis thaliana, 70 Archaeophytes, 5 Armillaria spp., 254, 255, 257, 258, 260, 262, 263, 265 A. borealis, 254 A. cepistipes, 254 A. gallica, 254 A. mellea, 25, 254, 255, 258, 263 A. ostoyae, 254, 258, 263 A. tabescens, 254, 263 Ascomycota, 215, 228 Aspergillus, 198, 199, 201, 207, 319, 320, 323, 324 A. alliaceus, 197 A. carbonarius, 197, 200–207 A. fumigatus, 326 A. niger, 197, 200–204 A. ochraceus, 197–201, 203, 205, 207 A. section Circumdati, 198–199 A. westerdijkiae, 197, 199–201, 203–206 laboratory diagnosis, 367–368 molecular detection, 368–373 Atopic dermatitis, 338, 339 B Bait tests, 60 Barcoding, 42, 72, 481 Basal fungal lineages, 215 Basidiobolus, Basidiobolaceae, 467, 468 Basidiomycota, 215, 217, 228 Index Batrachochytrium dendrobatidis, 13 Bayesian inference, 442, 456 Biosynthetic pathway 20-carbon PUFA biosynthesis, 476 n-3 PUFA biosynthetic pathway, 476 n-6 PUFA biosynthetic pathway, 475, 476 Blastocladiomycota, 215 Blastomyces dermatitidis laboratory diagnosis, 373–374 molecular detection, 374 Blood culture, 322, 323, 326, 327 whole, 317–328 Blood stream infection (BSI), 322, 327 therapy, 322 Blumeria graminis, 11 Botryosphaeria, 25 Buffon’s law, 7 C Calmodulin gene, 140 Cambrian, 467 Candida, 318–324 C. albicans, 318, 319, 323, 324, 326, 327 C. glabrata, 319, 321, 323, 326 C. krusei, 319, 321, 323, 326 C. parapsilosis, 319, 326 C. tropicalis, 319, 324, 326 laboratory diagnosis, 377–378 molecular detection, 378–382 Candidaemia, 318–319, 321 Carbon assimilation profiles, 479–480 Ceratocystis fagacearum, 13, 16 Cerotelium, 9 Characteristics barcode, 481 biochemical, 213 chemotaxonomical, 241 diagnostic importance, 230 ecological, 214 growth temperature, 471–473 isoenzyme patterns, 473 mating compatibility, 471, 473 mesophilic, 470, 472, 473 metabolic, 214 molecular, 214, 227, 461–481 morphological, 213, 215, 227, 230, 464, 467, 470–473, 477, 481 Index phylogenetic relevance, 221, 223, 237 physiological, 213, 241 synapomorphic, 471 thermophilic, 464, 472, 473 thermotolerant, 470, 471 zygospore, 468, 470, 471 Charcoal rot, 108, 181 Chasmothecial appendage, 85, 86, 88, 89 Chemotype 3ADON chemotype, 161–167, 169–174 15ADON chemotype, 161–167, 169–174 molecular chemotype, 164, 169, 170, 173 NIV chemotype, 160–162, 165–167, 170–172, 174 Chickpea, 80–83 China, 161, 163, 165, 166, 169, 171, 173 Chlamydoabsidia, 441, 442, 444, 445, 450 Chromatography, affinity, 323, 326–328 Chytridiomycota, 215 Chytrids, 468 Ciborinia camelliae, 11 Cicer arietinum, 80 Circinella, 447 Cladosporium, 7 C. subtilissimum, 8 Classification DNA-based, 467, 472, 480 morphology-based, 467, 472 traditional, 467 Cleaved amplified polymorphic sequence (CAPS), 147 Clinical specimens BAL, 371, 372, 378, 390 blood, 363, 368, 371, 372, 375, 377, 378, 381, 382, 386, 390, 391, 393, 394 CSF, 363, 385, 386 serum, 371, 372, 378, 382, 386, 391 serum, plasma, 363, 371 skin, 360, 373, 388, 391, 394, 395 tissue, 361, 367, 368, 371, 373, 378, 383, 384, 388, 390–392, 394 urine, 372, 386 Coccidioides spp. laboratory diagnosis, 383 molecular detection, 384 Cochliobolus carbonum, 9 Coleosporium helianthi, 10 491 Compatibility mating, 471 tests, 257 Confluent and reticulate papillomatosis, 338, 339 Coniophora puteana, 253, 261 cox2, 37, 38, 41 C-reactive protein (CRP), 321 Cronartium C. occidentale, 20 C. quercuum f.sp. fusiforme, 16 C. ribicola, 18, 19 Cryphonectria parasitica, 24 Cryptococcus neoformans laboratory diagnosis, 385–386 molecular detection, 386–387 Cultivation-dependent method, 278, 287 Culture collections, 214 Cunninghamella, 444, 445, 464, 471, 475–478 Cutinase gene, 141 Cytochrome c oxidase subunit I (COI), 146 D D-arabinitol, 321 D1/D2 domain, phylogenetic trees, 345, 348, 351 D-dimer, 321, 322 Dematiaceous laboratory diagnosis, 388 molecular detection, 388–389 Denaturing gradient gel electrophoresis (DGGE), 279, 281–282, 287 Deoxynivalenol (DON), 160 Derived cleaved amplified polymorphic sequence (dCAPS), 147, 148 Dermatophytes laboratory diagnosis, 396 molecular detection, 396–397 Devonian, 469 b-D-glucan, 320 Diagnostic tools, 184–187 Dictyosome, 467 Dikarya, 215 Dikaryomycota Ascomycota, 468, 480 Basidiomycota, 464, 469, 470, 480 492 Dimorphic, 360, 373, 382, 384, 390 Dimorphic ascospores, 83 Discula destructiva, 15 Divergence of Asco- and Basidiomycota, 469 of Metazoa and fungi, 469 Diversity ecological, 462, 464, 467 global biodiversity information faculty, 481 DNA, 303, 307–310, 317–328 array, 100, 137, 144 background, 326 extraction, 68, 165 fingerprinting, 279–282, 287, 289 human background, 326 hybridization, 307 manipulations, 362–364 methylation, 326 microarray, 324, 328 nuclear ribosomal, 303 polymorphism, 307 rDNA, 467–475, 477–478 ribosomal, 468, 473, 474, 477, 478, 481 DNA barcodes, 80, 89, 145–147, 149, 214, 215, 240, 242 alternative barcode markers, 215 barcode markers in combination, 234 beta-tubulin, 234 calmodulin, 234 Consortium for the Barcode of Life (CBOL), 214 cox1, 214–215 cytochrome b (cob), 215 IGS, 240 internal transcribed spacer (ITS), 215, 217–222, 224–228, 239–240 16S rDNA, 215 28S rDNA, 221, 235, 240 translation elongation factor 1 alpha, 221, 227, 234 universal barcode marker, 214 Donkioporia expansa, 253 E Ecological importance, 72 Electrophoresis fluorescent capillary, 477, 478 Index pulsed field gel electrophoresis (PFGE), 462, 477, 479 Endocommensal, 464 Endocronartium harknessii, 21 Endomycorrhiza, arbuscular, 468, 469 Endophyte, 277, 278, 280–282, 285, 288, 289 Endophytic fungi, 277–289 Endo-polygalacturonase gene (pgI), 140 Entoleuca mammata, 13 Entomophthoromycotina arbuscular endomycorrhiza, 468, 469 Entomophthorales, 464, 467, 468, 476, 484 Enzyme-linked immunosorbent assay (ELISA), 62, 168, 258 Erysiphales, 84 Erysiphe E. pisi, 85, 86 E. trifolii, 81, 85–89 Evolution actin, 462–464, 469, 471, 474, 475 EF-1 alpha, 469, 471, 473 linear, 469 tubulin, 462–464, 468 Exaptation, 14 F Fahrenholz’s rule, 11 Fatty acids n-3 PUFA, 476 n-6 PUFA, 475, 476 polyunsaturated, 474–476 PUFA, 474–476 Fennellomyces linderi, 447, 448, 457 Filamentous fungi, 195 Fingerprinting amplified fragment length polymorphism (AFLP), 474, 475, 479 microsatellite PCR, 477, 479 PCR-RFLP, 477–479 pulsed field gel electrophoresis (PFGE), 462, 479 restriction fragment length polymorphism (RFLP), 477, 479 whole-genome, 479 Finland, 159–174 Forestry, 252 Index Fruit bodies, 252, 256, 267, 268 Fumonisins biosynthesis, 120 effects, 109 F. nygamai, 108, 115 F. proliferatum, 108, 111, 114–118 F. verticillioides, 108, 111, 114–120 structure, 108 toxic action, 108–110 Fungi, Dikaryomycota, 470 Fusarium head blight (FHB), 159 Fusarium oxysporum formae speciales, 132, 134, 136–145, 147, 150 Fusarium spp., 93–102, 185, 319 biological species concept, 113, 115 F. asiaticum, 163, 167, 172 F. avenaceum, 161 F. cerealis, 159–174 F. culmorum, 159–174 F. graminearum, 159–174 F. ussurianum, 173 laboratory diagnosis, 389–390 molecular detection, 390 morphological species concept, 113–115 phylogenetic species concept, 113–116, 119 Fuscoporia torulosa, 266 G Ganoderma spp., 255, 256, 258, 263, 265 Gene, 302, 303, 308 Ac12RL3 gene, 206, 207 actin, 462–463, 469, 471, 474, 475 actin (act), 302, 304, 307 beta-tubulin (btub), 307, 308 cytochrome, 478–479 D5-desaturase, 476 D9-desaturase, 476 D12-desaturase, 476 encoding o3-desaturase, 476 GenBank, 465, 466, 481 lactate dehydrogenase B, 475 orthologous, 303 otapksPN gene, 207, 208 paralogous, 303 pks gene, 200, 207–208 protein coding, 307 493 repetitive, 303 single copy, 303, 307, 481 translation elongation factor (tef), 303, 307 translation elongation factor-1alpha, 474, 475 Genome analysis, 145 mitochondrial, 480 Mucor, 462, 464, 465, 471, 472, 476, 478–480 Phycomyces, 462, 466, 480, 481 Rhizopus, 462, 464, 466, 470, 472–476, 478–481 sequencing, 145–147 Genomes OnLine Database (GOLD), 145 Genomics, 68–71 Genotyping microsatellite PCR, 477, 479 PCR-RFLP, 477–479 pulsed field gel electrophoresis (PFGE), 462, 477, 479 randomly amplified polymorphic DNA (RAPD), 462, 473, 477, 479 restriction fragment length polymorphism (RFLP), 473, 477–479 whole-genome, 479 Geosiphonales, Geosiphon, 468 Germany, 163, 165, 166, 169–171 Gloeophyllum spp., 253 Gloeophyllum sepiarium, 262 Glomeromycota, 215 Glomeromycota, Geosiphonales, 467, 468 Glucoamylase, 474 Gongronella, 441, 442, 444, 445 Gougerot and Carteaud syndrome, 338, 340 gp43, 422–424, 427–430 Graminicolous downy mildews, 37, 39 Group I introns, 81–84, 89 Gymnosporangium fuscum, 23 H Halteromyces, 444, 445 Herbarium specimens, 42 Hesseltinella, 444 Heterobasidion annosum sensu lato (s.l.), 254, 255, 257, 260, 262–264, 267 H. abietinum, 255, 262, 264 494 H. annosum sensu stricto (s.s.), 255, 262–264 H. parviporum, 255, 262–264 High performance liquid chromatography (HPLC), 136 Histoplasma capsulatum laboratory diagnosis, 391 molecular detection, 391–392 Holocene, 3, 5 Homogocene, 3–27 Hyaloperonospora, 39 Hybridization probes, 187, 188 Hypertrophies, 53, 60 Hypocrea, 185 I Identification, 56, 62 of Aspergillus, 368–369, 372 barcode, 481 of Blastomyces dermatitidis, 373, 374 of Candida, 374, 376 of chromoblastomyces, 388–389 of Coccidioides, 383, 384 of Cryptococcus neoformans, 387 of dermatophytes, 396, 397 of filamentous fungi, 372 of Fusarium, 389, 390 of Histoplasma, 390, 391 of molds and yeasts, 359 morphological, 464, 472, 473, 477, 481 phenotype-based identification, 397 physiochemical, 472, 473 preliminary identification, 377 species identification, 359, 381, 396, 397 of Trichosporon, 393 of yeasts, 359 of Zygomycetes, 394 Immunological test methods, 62 Infection fungal, 471, 476–478, 480 invasive fungal infections (IFI), 319–321, 323 nosocomial, 319 zygomycosis, 472, 476, 478, 480 Inonotus spp., 254, 263, 265 I. tomentosus, 255 Intergenic spacer (IGS), 138–139 Index Internal transcribed spacer (ITS), 41–44, 80–83, 85–89, 138–139 Inter-simple sequence repeat (ISSR), 137, 279–281 Intraspecific phylogeny, 58 Invasive/disseminated fungal infections (IFI) angioinvasive moulds, 389 angioinvasive zygomycosis, 393 cryptococcosis, 385 diagnosis of IFI, 361–362 diagnosis of invasive aspergillosis, 368, 371 invasive aspergillosis, 367, 389 invasive candidiasis, 375 invasive candidosis, 382 invasive fusariosis, 389 invasive mold infection, 372, 389 invasive pulmonary aspergillosis, 362, 371 invasive trichosporonosis, 392 limited invasive disease, 367 progressive invasive disease, 367 risk factors for invasive aspergillosis, 367 Isozyme, 134, 136 analysis, 39, 257, 307 iSSR, 41 ITS regions, phylogenetic trees, 345, 348, 351 K Kickxellomycotina Harpellales, 464, 467, 468 Kickxellales, 464, 467, 468 Kretzschmaria deusta, 254 Kuehneola, 9 L Laetiporus sulphureus, 254 Large-subunit, D2, 477 Lens culinaris, 80 Lentamyces L. parricida, 446–449, 452, 457 L. zychae, 446–449, 452, 457 morphological aspects, 447, 448 mycoparasitism, 447 physiological aspects, 447 Index RFLP, in silico, 448 sucker-like substrate mycelium, 447 Lentil, 80, 81, 84–89 Lichtheimia discrimination from Absidia, 453 L. corymbifera, 445, 446, 450 L. hyalospora, 446, 450 L. ornata, 446, 450 L. ramosa, 446, 450, 454, 457 L. sphaerocystis, 446, 450 morphological aspects, 445 physiological aspects, 445 Lichtheimiaceae, 440, 445–446, 450 Lichtheimiaceae, Lichtheimia, 471 Life cycle, 51–73 Ligniera, 71 Loop-mediated isothermal amplification (LAMP) methods, 417–435 LOOXSTER1, 327 Lophodermium pinastri, 22 LSU, large-subunit, 477 Lysis mechanical, 325, 326 pathogen cells, 325 M Macrophomina phaseolina biochemical and serological characterization, 182 clasification and nomenclature, 180–181 identification and characterization, 181 morphological and cultural characteristics, 181 Maize, mycotoxins, 108, 111 Malassezia spp identification, 337, 338, 340, 341, 344, 345 M. caprae, 343 M. dermatis, 341–343, 348 M. equi, 343 M. equina, 343 M. furfur, 338, 340–342, 345, 348 M. globosa, 338, 339, 341–343, 345, 348 M. japonica, 342, 343, 348 M. nana, 342, 343 molecular techniques, 338, 343 M. pachydermatis, 340–343, 345, 348, 351 495 M. restricta, 339, 341, 342, 345, 348 M. sloffiae, 345 M. sympodialis, 338, 341–343, 348, 351 M. yamatoensis, 342, 343, 348 phenotypical and physiological features, 341 rDNA genes, 344 Melampsora M. hypericorum, 9, 10 M. larici-populina, 21 M. medusae, 20, 21 M. occidentalis, 20 Melampsora x columbiana, 21 Meruliporia incrassata, 253 Mesomycetozoea, Eccrinales, 464 Metagenomics, 148, 149 Metazoa, 469 Microarrays, 70, 142, 144–145, 147–150, 268, 269, 461 Microcyclus ulei, 26 Minimum ages, 469, 470 Molecular actin, 462–464, 469, 474, 475 chemotype, 164, 169, 170, 173 clocks, 469 data available, 71 detection, 131–150 lactate dehydrogenase B, 474, 475 tools, restriction fragment length polymorphism, 183–184 translation elongation factor-1alpha, 462–464, 474, 475 Molecular assays, amplicon size, PCR and sequencing fluorescence-based PCR, 395 FRET probes, 395 in-house and commercial PCR, 372 LAMP, 380 multiplex PCR, 372, 373, 387, 390, 394 nested PCR, 374, 378, 384, 386, 387, 393 panfungal PCR, 367, 368 PCR and cryptococcosi, 386 PCR direct on BAL, 371 PCR ELISA, 368, 371 PCR-RFLP, 387, 396, 397 PCR serotyping, 387 PCR vs. blood cultures, 382 qPCR, 371, 395 496 quantitative real time, 371, 393 real time, 371–373, 384, 393, 395 rep-PCR, 360, 372, 374, 381, 384, 392 seminested PCR, 391, 394 Molecular basis of plasmodiophorid infection, 69 Monitoring, 60 Monophyletic, 39 origin, 132 Morphological characters, 36 Morphology subsporangial swelling, 471 trophocyst, 471 Morphotypes, 280, 284–286, 288 Mortierellomycotinained Endogonales, 469 Mortierellales, 469 Mucor, 213–243 M. circinelloides, 216, 221, 222, 224, 226, 229–231, 233, 236, 238–240, 447, 448, 457 M. circinelloides f. circinelloides, 222, 227, 233, 238, 239 M. circinelloides f. griseo-cyanus, 221, 233, 234, 238–240 M. circinelloides f. janssenii, 236 M. circinelloides f. lusitanicus, 227, 228, 233, 236, 238, 239 M. corymbifer, 445 M. mucedo, 226, 236 M. racemosus, 221, 226, 236 polyphyletic, 235–237 Mucorales Absidia, 470–472, 476, 478 Actinomucor, 471 Cunninghamella, 471, 476, 478 facultative parasites, 439, 447, 452 Lichtheimiaceae, 471 Mucor, 471, 472, 476, 478 mycoparasitic species, 440, 446 opportunistic pathogens, 439 Pilaira, 471–472 Pilobolaceae, 471 Rhizomucor, 472–473, 476, 478 Rhizopus, 470, 472–474, 476, 478 saprobes, 439 Umbelopsidaceae, 469 Mucormycoses, 216, 235, 236, 240, 439, 445 Index Mucoromycotina Mucorales, 469, 476 Multilocus genotyping (MLGT), 161, 163, 167–168, 170, 172 Multiplex PCR, 141 Multiplex polymerase chain reaction (m-PCR), 203–204 Mycocladaceae, 446 Mycocladiaceae, 470 Mycocladus, 470–471 Mycocladus verticillatus, 446 Mycorrhiza ectomycorrhiza, 464 endomycorrhiza, 469 Mycosphaerella M. fijiensis, 4 M. musicola, 4 M. populicola, 22 Mycotoxins, 195–198, 207–208 N Necrosis, 169, 171 Neocallimastigomycota, 215 Neolithic, 3, 5, 6, 17 Neophytes, 5 Next-generation sequencing, 147 Nivalenol (NIV), 160 NIV chemotype, 160–162, 165–167, 170–172, 174 Non-sporulating endophytic fungi, 284–285, 289 Nucleic acid amplification technique (NAT), 323–328 Nucleic-acid-based detection methods, 63–68 O Obligate intracellular parasites, 52 Ochratoxigenic fungi, 195–209 Ochratoxin A (OTA) biosynthetic pathway genes, 206–208 chemical structure, 196 effects, 196–197 molecular marker, 197–198 PCR detection and quantification, 198–208 producer, 197 Index regulations, 197 Ochroconis gallopava, 417–435 Oligonucleotide, 303 microsatellite, 477 universal, 303 Oligoporus placenta, 253 On-site PCR, 68 Oospore ornamentation, 39 Ophiostoma, 16 O. novo-ulmi, 17 O. ulmi, 17 Opportunistic, 358, 360, 361, 367, 373, 392 Origin of Mortierellales, 469 of Mucorales, 469 P Paleozoic, 470 Panfungal assays, 364–367 Paracoccidioides P. brasiliensis, 418, 419, 421–431 P. lutzii, 423, 424 Paracoccidioidomycosis, 417–435 Parasites obligate, 464 Pathogen, 470, 472, 476–479 Pathogenicity, 132, 136, 149, 150, 163, 168–169, 171–172 Pathogen release hypothesis, 8–10 Pathway, biosynthetic pathway, 476 Patients immunocompromised, 476 immunosupprimised, 476 PCR chemotyping, 165–167 PCR-fingerprinting techniques ITS, 183, 185–187 ITS-RFLP, 186 Penicillium, 199, 204 P. nordicum, 200, 204, 207, 208 P. verrucosum, 197, 200, 204, 208 Perenniporia fraxinea, 254 Peridiopsora, 9 Peronospora P. farinosa, 37 P. sparsa, 43 Phacidium infestans, 22 Phagomyxa, 57 497 Phagomyxids, 52, 57, 71, 72 Phakopsora, 9 Phakopsora P. meibomiae, 9 P. pachyrhizi, 9 Phellinus spp., 254, 255, 263, 265 P. noxius, 263 P. sulphurascens, 261 P. weirii, 255 Phenotypic characters, 39 Phlebia spp., 255, 263 Phylogenetic analysis, parsimony analysis, 87 Phylogenetic marker actin (act), 442–444, 456 ribosomal DNA, 442, 444, 448, 449, 453–456 translation elongation factor 1 alpha (tef), 442–444, 456 Phylogenetics, 462, 465–466, 469, 472, 481 coherence, 470 molecular, 462, 470, 481 multigene, 471 Phylogeny actin, 462–464 DNA-based, 466–468, 470, 472, 473, 480, 481 Fungal Tree of Life, 481 molecular, 466, 472, 473 multigene, 471 phylogenetic analyses, 360 phylogenetic classification, 360 phylogenetic investigation, 358 phylogenetic markers, 398 phylogenetic relationships, 396 phylogenetic studies, 358 protein, 462–466, 468, 481 rpb1, 462, 469 rpb2, 462 Phytomyxea, 56, 57 Phytophthora, 4, 15, 25, 40, 44 P. cinnamomi, 25 P. infestans, 4 P. lateralis, 15 Pilaira, 471–472 Pink ear rot F. proliferatum, 111, 116, 117 F. subglutinans, 111, 116, 117 498 F. verticillioides, 111, 116, 117 infection, 111 inoculum, 111 Pityriasis capitis, 338 Pityriasis versicolor, 338, 339 Plant diseases, 59, 63 Plant Pathogen Barcode, 146 Plasmodiophora brassicae, 52, 55–57, 59–63, 70–72 Plasmodiophorids, 51–73 Plasmopara P. halstedii, 37, 40–44 P. viticola, 41, 43 Podosphaera leucotricha, 23 Polygalacturonase, 474–475 Polymerase chain reaction (PCR), 80, 82, 83, 86, 136–143, 145, 147, 148, 150, 258–259, 323–325, 327, 419, 421, 427–431, 433, 477–480 amplified fragment length polymorphism (AFLP), 479 DNA sequencing, 266–267 group specific, 98–100 microsatellite PCR, 479 multiplex, 317–328 nested PCR, 262, 267 PCR-RFLP, 477–479 qPCR, 326, 327 random amplified polymorphic DNA (RAPD), 259–261, 268 random amplified satellites (RAMS), 263 RAPD, 477, 479 real-time, 324, 326, 477–479 real-time PCR, 266 restriction fragment length polymorphism (RFLP), 260–261 sequence specific oligonucleotide probe (SSOP), 269 species specific, 96–98 taxon-specific PCR, 261–269 terminal restriction fragment length polymorphism (T-RFLP), 261 Polymerase chain reaction (PCR) assay, 198–209 Polymyxa, 57, 58, 62 P. betae, 57–59, 62, 63, 67 P. graminis, 57–59, 62, 63, 69–71 Polyphyly, 464 Index Population structure, 137 Post harvest diseases, 240 Powdery mildew, 84–86, 89 Powdery scab, 59, 61 Primer, 303 universal, 303, 307 Proabsidia, 441, 442 Procalcitonin (PCT), 321 Projects, genome, 465–466, 480–481 Proteome, 70 Protoabsidia, 441, 442 Pseudoabsidia, 441, 442, 446 Psoriasis, 338, 340 Public databases, 214, 235 Puccinia P. carthami, 26 P. helianthi, 10 P. irrequiseta, 9 P. jaceae var. diffusa, 9 P. psidii, 11, 18, 21 P. tanaceti, 9 Pucciniastrum corni, 15 Pythiales, 38, 39 Q qPCR, 161 Quantitative real time PCR, 205 R Radiation basidiomycete radiation, 469 pezizomycotina radiation, 469 Random amplified polymorphic DNA (RAPD), 40, 133, 134, 136–137, 198, 200, 202–204, 228, 279–281 Real-time PCR, 139, 142–144, 148, 150, 204–205, 207 Real -time PCR technology SYBR Green, 188 TaqMan, 188 Red ear rot colonization, 112, 113 DON, 112, 113 F. acuminatum, 111 F. avenaceum, 111 F. chlamydosporum, 111 F. culmorum, 111, 113, 116–119 Index F. equiseti, 111 F. graminearum, 111–113, 117–119 F. heterosporum, 111 F. poae, 111 F. semitectum, 111 F. sporotrichioides, 121 infection, 112 inoculum, 112, 113 symptoms, 109 Reference material, 234 Relationships, 467, 469, 470, 472, 473, 481 Relevance clinical, 461 Resting spores, 55, 56, 60, 62, 67–69 Restriction fragment length polymorphism (RFLP), 39, 133–136, 279–281, 287, 307 Rhizina undulata, 25 Rhizomorphs, 254 Rhizomucor, 472–473, 475, 476, 478, 479 Rhizopus, 213–243 R. americanus, 235 R. arrhizus, 216, 224, 226, 228, 230, 232, 235–238, 240, 241 R. arrhizus var. arrhizus, 221, 228, 236, 241 R. arrhizus var. delemar, 241 R. caespitosus, 227, 232, 238 R. homothallicus, 227, 232, 235, 238 R. lyococcus, 235, 236 R. microsporus, 216, 232, 235, 238 R. microsporus var. chinensis, 226 R. microsporus var. microsporus, 227, 235 R. microsporus var. oligosporus, 226 R. microsporus var. rhizopodiformis, 226 R. schipperae, 227, 238 R. sexualis, 227, 232, 235, 238 R. stolonifer, 224, 226, 228, 230, 232, 235–238 R. stolonifer var. americanus, 235, 237 R. stolonifer var. lyococcus, 235 R. stolonifer var. sexualis, 235 R. stolonifer var. stolonifer, 226, 235 Rhynchosporium secalis, 13 Rhytisma americanum, 22 Ribosomal DNA (rDNA), 303, 307 cluster, 303, 307 499 intergenic spacer (IGS), 307 ITS, 469, 471–475, 478 large subunit (LSU), 303, 307 LSU, 469, 472, 473, 477–478 18S, 307, 308 28S, 307, 308 small subunit (SSU), 307 SSU, 468, 469, 472, 473, 477–478 variability, 307 Rosellinia necatrix, 255 S Saprobe, 464, 473 Sclerotinia S. sclerotiorum, 81–84, 89 S. trifoliorum, 80–84, 89 Seborrhoeic dermatitis, 338, 339 Section Discolor chemotypes, 116 F. culmorum, 115 F. graminearum, 115, 116 F. pseudograminearum, 116 morphology, 116 Section Liseola F. anthophilum, 114 F. moniliforme, 114 F. proliferatum, 114, 115 F. subglutinans, 114, 115 mating populations, 114, 115 phylogenetic analyses, 122 Seiridium, 16 Sepsis, 326, 327 causative pathogen, 317–319, 326 therapy, 319–322 Sequence-based classification, 41 Sequence characterized amplified region (SCAR), 138, 140 Sequence tagged sites (STS), 138 454 Sequencing, 147 Sequencing, direct, 477, 480 Serpula lacrymans, 13, 253 Siepmannia RFLP, restriction patterns, 447, 448 S. lariceti, 447, 448, 457 S. parricida, 447, 449 S. pineti, 447, 457 S. zychae, 447, 449 500 Simple sequence repeats (SSRs), 137–138, 279–281 Single nucleotide polymorphic sites (SNPs), 82 Single-nucleotide polymorphisms (SNPs), 42, 139, 143, 147, 148 Sirococcus clavigignenti-juglandacearum, 15 SNPs. See Single nucleotide polymorphism Sodium dodecyl sulfate polyacrilamide gel electrophoresis (SDS-PAGE), 257 Solexa, 147 SOLiD, 147, 149 Sorosphaera, 53 S. veronicae, 57, 61, 69 S. viticola, 61, 69 Species specific PCR calmodulin, 115, 117 F. culmorum, 116–118, 121, 122 F. graminearum, 110, 116–119, 121 F. proliferatum, 108, 111, 116–118 F. subglutinans, 115–118 F. verticillioides, 108, 116–120 IGS, 117–119 ITS, 117, 118 Species-specific primers, 63, 163 specimens, unknown, 214 Sphaeropsis sapinea, 13 Spongospora subterranea, 57, 59–62, 69 18S rDNA, 37, 38, 41 28S rDNA, 37, 41 Stegophora ulmea, 22 Stem rot, 80–84 Sterile mycelia, 284, 285 Straminipila, 36 Suspensor, appendages, 470 Systematics Index Fungorum, 481 Index Fungorum of CABI Bioscience, 471–472 molecular, 462, 473 MycoBank, 481 Index Technology, sequencing, 481 Tens rule, 6, 22–24 Tieghemella, 441, 442 Timber, 251–269 Toxin-specific PCR EF-1a, 119 F. culmorum, 121 F. graminearum, 121 F. sporotrichioides, 110, 121 FUM cluster, 120 fumonisins, 108–109, 119–121 F. verticillioides, 108, 119, 120 IGS-RFLP, 119 TRI cluster, 121 Transcriptome analysis, 70 Transformation, Agrobacterium tumefaciens-mediated, 481 Translation elongation factor-1a gene (TEF-la), 137 Transposons, 139–140 Tree stability, 252–254, 263, 265 Tri7, 161–165, 169–170, 172, 173 Tri13, 161, 162, 164, 165, 169–170, 172, 173 Trichoderma, 185 Trichomycetes Asellariales, 464, 467, 468 Trichophyton rubrum, 13 Trichosporon spp. laboratory diagnosis, 392–393 molecular detection, 393 Trichothecenes DON, 109, 110, 121 F. acuminatum, 110 F. culmorum, 110, 121 F. equiseti, 110 F. graminearum, 110, 121 F. sporotrichioides, 110, 121 structure, 109, 110 T-2, 109, 110 toxic action, 110 Tubulin, FtsZ, 464 T Taphrina, 14 Taxonomic concept, 36 Taxonomic position, 56 Taxonomy, 358, 360, 392, 395, 396 U Universal primers, 462, 474 Uredinales, 11 Uromyces heterogeneus, 23 Index 501 V Z Vegetative compatibility groups (VCGs), 132–134, 136 Venturia, 10, 14 V. inaequalis, 13, 23 V. inopina, 10 V. populina, 10 Verticillium, 185 Virus transmission, 59, 61–63, 67 VYOO1, 325, 326 Zearalenone (ZEN), 160 Zoopagomycotina Dimargaritales, 464 Zoopagales, 464 Zygomycetes, 215–217, 221–224, 231, 234, 241, 464, 468, 474, 478, 481 laboratory diagnosis, 394 molecular detection, 394–395 Zygomycetous fungi, 461–481 Zygomycosis antimycotics, 472 entomophthoromycosis, 476 mucormycosis, 473, 481 Zygomycota, 215, 221, 228, 230 Basidiobolales, 468 Entomophthoromycotina, 462, 467, 468 Kickxellomycotina, 462, 467, 468 Mortierellales, 469 Mortierellomycotinained, 469 Mucoromycotina, 462 Zoopagomycotina, 462, 467 Zygospore, 468, 470, 471 W White blister rusts, 37, 39, 44 Wood rotting fungi, 251–269 brown rot, 251, 253, 254, 256 butt rot, 252, 254–256, 263 indoor wood decay fungi, 252–253, 257, 258 root rot, 254, 255, 263, 267 white rot, 251, 253, 254 X Xylanase-3 gene, 140