Fungal Diseases - An Emerging Threat To Human
Fungal Diseases - An Emerging Threat To Human
Fungal Diseases - An Emerging Threat To Human
Workshop Summary
NOTICE: The project that is the subject of this report was approved by the Governing
Board of the National Research Council, whose members are drawn from the councils of
the National Academy of Sciences, the National Academy of Engineering, and the Institute
of Medicine.
Financial support for this project was provided by the U.S. Department of Health and Hu-
man Services: National Institutes of Health, National Institute of Allergy and Infectious
Diseases, Centers for Disease Control and Prevention, Food and Drug Administration,
and the Fogarty International Center; U.S. Department of Defense, Department of the
Army: Global Emerging Infections Surveillance and Response System, Medical Research
and Materiel Command, and the Defense Threat Reduction Agency; U.S. Department of
Veterans Affairs; U.S. Department of Homeland Security; U.S. Agency for International
Development; American Society for Microbiology; sanofi pasteur; Burroughs Wellcome
Fund; Pfizer, Inc.; GlaxoSmithKline; Infectious Diseases Society of America; and the
Merck Company Foundation. Any opinions, findings, conclusions, or recommendations
expressed in this publication are those of the author(s) and do not necessarily reflect the
view of the organizations or agencies that provided support for this project.
International Standard Book Number-13: 978-0-309-21226-7
International Standard Book Number-10: 0-309-21226-X
Additional copies of this report are available from the National Academies Press, 500 Fifth
Street, N.W., Lockbox 285, Washington, DC 20055; (800) 624-6242 or (202) 334-3313
(in the Washington metropolitan area); Internet, http://www.nap.edu.
For more information about the Institute of Medicine, visit the IOM home page at: www.
iom.edu.
Copyright 2011 by the National Academy of Sciences. All rights reserved.
Printed in the United States of America
The serpent has been a symbol of long life, healing, and knowledge among almost all
cultures and religions since the beginning of recorded history. The serpent adopted as a
logotype by the Institute of Medicine is a relief carving from ancient Greece, now held by
the Staatliche Museen in Berlin.
Cover images: Front (upper): Little brown bats with white-nose syndrome, New York,
photo courtesy of New York Department of Environmental Conservation; Front (lower):
Yellow stripe rust on wheat, photo courtesy of Stephen A. Harrison, Louisiana State Uni-
versity Agricultural Center. Spine: The Panamanian golden frog (Atelopus zeteki), photo
courtesy of Wikimedia Commons, photo by Brian Gratwicke, Smithsonian Conservation
Biology Institute. Back: Geomyces destructans, shown in a false-color SEM image (fungus
hyphae are yellow, green, and orange; spores are blue), image reprinted from Chaturvedi
et al. (2010) Morphological and Molecular Characterizations of Psychrophillic Fungus
Geomyces destructans from New York Bats with White Nose Syndrome (WNS). PLoS
ONE 5(5): e10783. Doi: 10.1371/journal.pone.0010783.
Suggested citation: IOM (Institute of Medicine). 2011. Fungal Diseases: An Emerging
Threat to Human, Animal, and Plant Health. Washington, DC: The National Academies
Press.
“Knowing is not enough; we must apply.
Willing is not enough; we must do.”
—Goethe
The National Academy of Engineering was established in 1964, under the charter of
the National Academy of Sciences, as a parallel organization of outstanding engineers.
It is autonomous in its administration and in the selection of its members, sharing with
the National Academy of Sciences the responsibility for advising the federal government.
The National Academy of Engineering also sponsors engineering programs aimed at
meeting national needs, encourages education and research, and recognizes the superior
achievements of engineers. Dr. Charles M. Vest is president of the National Academy of
Engineering.
The Institute of Medicine was established in 1970 by the National Academy of Sciences
to secure the services of eminent members of appropriate professions in the examina-
tion of policy matters pertaining to the health of the public. The Institute acts under the
responsibility given to the National Academy of Sciences by its congressional charter to
be an adviser to the federal government and, upon its own initiative, to identify issues of
medical care, research, and education. Dr. Harvey V. Fineberg is president of the Institute
of Medicine.
The National Research Council was organized by the National Academy of Sciences in
1916 to associate the broad community of science and technology with the Academy’s
purposes of furthering knowledge and advising the federal government. Functioning in
accordance with general policies determined by the Academy, the Council has become
the principal operating agency of both the National Academy of Sciences and the Na-
tional Academy of Engineering in providing services to the government, the public, and
the scientific and engineering communities. The Council is administered jointly by both
Academies and the Institute of Medicine. Dr. Ralph J. Cicerone and Dr. Charles M. Vest
are chair and vice chair, respectively, of the National Research Council.
www.national-academies.org
FORUM ON MICROBIAL THREATS1
DAVID A. RELMAN (Chair), Stanford University and Veterans Affairs Palo
Alto Health Care System, Palo Alto, California
JAMES M. HUGHES (Vice-Chair), Global Infectious Diseases Program,
Emory University, Atlanta, Georgia
LONNIE J. KING (Vice-Chair), Ohio State University, Columbus
KEVIN ANDERSON, Department of Homeland Security, Washington, DC
RUTH L. BERKELMAN, Center for Public Health Preparedness and
Research, Rollins School of Public Health, Emory University, Atlanta,
Georgia
DAVID BLAZES,2 Armed Forces Health Surveillance Center, Division of
Global Emerging Infectious Surveillance, Silver Spring, Maryland
ENRIQUETA C. BOND, Burroughs Wellcome Fund (Emeritus), Marshall,
Virginia
ROGER BREEZE, Lawrence Livermore National Laboratory, Livermore,
California
STEVEN J. BRICKNER,3 SJ Brickner Consulting, LLC, Ledyard,
Connecticut
PAULA R. BRYANT, Defense Threat Reduction Agency, Medical S&T
Division, Fort Belvoir, Virginia
JOHN E. BURRIS, Burroughs Wellcome Fund, Research Triangle Park, North
Carolina
ARTURO CASADEVALL,2 Albert Einstein College of Medicine, Bronx,
New York
PETER DASZAK, EcoHealth Alliance, New York, New York
JEFFREY S. DUCHIN, Public Health–Seattle and King County, Seattle,
Washington
JONATHAN EISEN, Genome Center, University of California, Davis
MARK B. FEINBERG, Merck Vaccine Division, Merck & Co., West Point,
Pennsylvania
JACQUELINE FLETCHER, Oklahoma State University, Stillwater
S. ELIZABETH GEORGE,3 Department of Homeland Security,
Washington, DC
JESSE L. GOODMAN, Food and Drug Administration, Rockville, Maryland
EDUARDO GOTUZZO, Instituto de Medicina Tropical–Alexander von
Humbolt, Universidad Peruana Cayetano Heredia, Lima, Peru
CAROLE A. HEILMAN, National Institute of Allergy and Infectious
Diseases, National Institutes of Health, Bethesda, Maryland
1╛╛Institute of Medicine Forums and Roundtables do not issue, review, or approve individual docu-
ments. The responsibility for the published workshop summary rests with the workshop rapporteurs
and the institution.
2╛╛Forum member since September 1, 2011.
3╛╛Forum member until December 31, 2010.
v
DAVID L. HEYMANN, Health Protection Agency, London, United Kingdom
PHILIP HOSBACH, sanofi pasteur, Swiftwater, Pennsylvania
STEPHEN ALBERT JOHNSTON, Arizona BioDesign Institute, Arizona
State University, Tempe
KENT KESTER, Walter Reed Army Institute of Research, Silver Spring,
Maryland
GERALD T. KEUSCH, Boston University School of Medicine and Boston
University School of Public Health, Boston, Massachusetts
RIMA F. KHABBAZ, Centers for Disease Control and Prevention, Atlanta,
Georgia
STANLEY M. LEMON, School of Medicine, University of North Carolina,
Chapel Hill
EDWARD McSWEEGAN, National Institute of Allergy and Infectious
Diseases, National Institutes of Health, Bethesda, Maryland
MARK A. MILLER, Fogarty International Center, Bethesda, Maryland
PAUL F. MILLER,4 Pfizer, Inc., Groton, Connecticut
STEPHEN S. MORSE,5 Center for Public Health Preparedness, Columbia
University, New York, New York
GEORGE POSTE, Complex Adaptive Systems Initiative, Arizona State
University, Tempe, Arizona
JOHN C. POTTAGE, JR., ViiV Healthcare, Collegeville, Pennsylvania
DAVID RIZZO,6 Department of Plant Pathology, University of California,
Davis
GARY A. ROSELLE, Veterans Health Administration, Department of Veterans
Affairs, Cincinnati, Ohio
ALAN S. RUDOLPH, Defense Threat Reduction Agency, Fort Belvoir,
Virginia
KEVIN RUSSELL, Armed Forces Health Surveillance Center, Department of
Defense, Silver Spring, Maryland
JANET SHOEMAKER, American Society for Microbiology, Washington, DC
P. FREDERICK SPARLING, University of North Carolina, Chapel Hill,
North Carolina
TERENCE TAYLOR, International Council for the Life Sciences, Arlington,
Virginia
MURRAY TROSTLE, U.S. Agency for International Development,
Washington, DC
MARY E. WILSON, Harvard School of Public Health, Harvard University,
Boston, Massachusetts
vi
Staff
EILEEN CHOFFNES, Director
LEIGHANNE OLSEN, Program Officer
KATHERINE McCLURE, Senior Program Associate
COLLIN WEINBERGER, Research Associate (until May 2011)
REBEKAH HUTTON, Research Associate (from June 2011)
ROBERT GASIOR, Senior Program Assistant (until March 2011)
PAMELA BERTELSON, Senior Program Assistant (since September 2011)
vii
BOARD ON GLOBAL HEALTH1
Richard Guerrant (Chair), Thomas H. Hunter Professor of International
Medicine and Director, Center for Global Health, University of Virginia
School of Medicine, Charlottesville
Jo Ivey Boufford (IOM Foreign Secretary), President, New York Academy of
Medicine, New York
Claire V. Broome, Adjunct Professor, Division of Global Health, Rollins
School of Public Health, Emory University, Atlanta, Georgia
Jacquelyn C. Campbell, Anna D. Wolf Chair, and Professor, Johns Hopkins
University School of Nursing, Baltimore, Maryland
Thomas J. Coates, Professor, David Geffen School of Medicine, University of
California, Los Angeles, Los Angeles, California
Gary Darmstadt, Director, Family Health Division, Global Health Program,
Bill & Melinda Gates Foundation, Seattle, Washington
Valentin Fuster, Director, Wiener Cardiovascular Institute Kravis
Cardiovascular Health Center Professor, Cardiology, Mount Sinai School
of Medicine, Mount Sinai Medical Center, New York, New York
James Hospedales, Coordinator, Chronic Disease Project, Health Surveillance
and Disease Management Area, Pan American Health Organization/World
Health Organization, Washington, DC
Peter J. Hotez, Professor and Chair, Department of Microbiology,
Immunology, and Tropical Medicine, The George Washington University,
Washington, DC
Clarion Johnson, Global Medical Director, Medicine and Occupational
Medicine Department, Exxon Mobil, Fairfax, Virginia
Fitzhugh Mullan, Professor, Department of Health Policy, George Washington
University, Washington, DC
Guy Palmer, Regents Professor of Pathology and Infectious Diseases, Director
of the School for Global Animal Health, Washington State University
Jennifer Prah-Ruger, Associate Professor, Division of Health Policy and
Administration, Yale University School of Public Health, New Haven,
Connecticut
Staff
Patrick Kelley, Director
Angela Mensah, Program Associate
1╛╛Institute of Medicine boards do not review or approve individual workshop summaries. The
responsibility for the content of the workshop summary rests with the authors and the institution.
viii
Reviewers
This report has been reviewed in draft form by individuals chosen for their
diverse perspectives and technical expertise, in accordance with procedures ap-
proved by the National Research Council’s Report Review Committee. The pur-
pose of this independent review is to provide candid and critical comments that
will assist the institution in making its published report as sound as possible and
to ensure that the report meets institutional standards for objectivity, evidence,
and responsiveness to the study charge. The review comments and draft manu-
script remain confidential to protect the integrity of the process. We wish to thank
the following individuals for their review of this report:
Although the reviewers listed above have provided many constructive com-
ments and suggestions, they were not asked to endorse the final draft of the report
before its release. The review of this report was overseen by Dr. Melvin Worth.
Appointed by the Institute of Medicine, he was responsible for making certain
that an independent examination of this report was carried out in accordance with
institutional procedures and that all review comments were carefully considered.
Responsibility for the final content of this report rests entirely with the authoring
committee and the institution.
ix
Acknowledgments
xi
xii ACKNOWLEDGMENTS
1╛╛Institute of Medicine (IOM) planning committees are solely responsible for organizing the work-
shop, identifying topics, and choosing speakers. The responsibility for the published workshop sum-
mary rests with the workshop rapporteurs and the institution.
2╛╛Sponsor as of October 1, 2010.
Contents
Workshop Overview 1
Workshop Overview References, 84
Appendixes
xiii
xiv CONTENTS
B Agenda 403
C Acronyms 409
D Glossary 413
E Forum Member Biographies 427
F Speaker Biographies 455
Tables, Figures, and Boxes
TABLES
WO-1 Number of Individual Animals Traded by the United States (2000–
2006), 21
WO-2 Disease Types and Associated Symptoms Caused by P. ramorum, 56
A6-1 Growth Tolerances for Fungi from Soils, Animals, and Plants at 2
Temperatures, 184
A14-1 Selected Papers Illustrating the Effects of Climate and Global Change
Factors on Specific Pathogen–Host Systems, 280
xvii
xviii TABLES, FIGURES, AND BOXES
FIGURES
WO-1 The fungal kingdom, 5
WO-1-1 Leafcutter ants tending their fungal garden, 10
WO-2 Diversity of fungal morphology, 6
WO-3 Depiction of starving Irish children in 1847 potato famine, 13
WO-4 The epidemiological triad, 16
WO-5 Global aviation network, 20
WO-6 Selected dispersal events of fungal pathogens, 22
WO-7 Environmental disturbances and dust storms contribute to the dispersal
of fungal spores, 24
WO-8 Change in precipitation between the 1971–2000 average and the
2091–2100 average in inches of liquid water/year, 27
WO-9 Incidence of systemic fungal disease has increased since the 1950s, 30
WO-10 Damage response framework, 31
WO-11 Microbial flora as a host defense, 33
WO-12 Map of the Pacific Northwest, comprising parts of British Columbia,
Canada, and the states of Washington and Oregon in the United States,
showing human and veterinary Cryptococcus gattii cases, 36
WO-13 Environmental sampling for Cryptococcus gattii in British Columbia
(2001–2009), 40
WO-14 Signs of bat white-nose syndrome (WNS), 42
WO-15 Spread of bat white-nose syndrome (WNS) in North America as of
April 21, 2011, 44
WO-16 Species affected by bat white-nose syndrome (WNS), 45
WO-17 Global distribution of Bd, 48
WO-18 A chytridiomycosis outbreak in southern mountain yellow-legged
frogs, 50
WO-19 Sudden oak death and ramorum blight, 55
WO-20 P. ramorum “migration” pathways, 57
WO-21 Wheat production regions worldwide, 59
TABLES, FIGURES, AND BOXES xix
A3-1 Fungal phyla and approximate number of species in each group, 142
A3-2 Lemonniera sp., 144
A3-3 The aero-aquatic ascomycete Helicoon gigantisporum produces
distinctive tightly coiled conidia, 144
A3-4 The smut Testicularia sp. develops in the ovary of grasses and (as
shown here) sedges, 144
A3-5 Perithecia of Pyxidiophora sp. (Laboulbeniomycetes) developed
in moist chamber on moose dung from Meredith Station, New
Brunswick, Canada, 144
A3-6 The ca. 8 cm wide basidiomata of Pycnoporus sp., a wide-ranging,
brightly colored, wood-decaying polypore, photographed at Barro
Colorado Island, Panama, 144
xx TABLES, FIGURES, AND BOXES
A10-1 The C. gattii outbreak expanded into, and emerged within, the United
States, 229
A10-2 Cryptococcus pathogenic species complex, 230
A10-3 Cryptococcus neoformans can reproduce unisexually and
bisexually, 235
A10-4 Sexual reproduction and the origin of an outbreak, 240
A14-1 The increase in goods (109 tons × km) moved in the United Kingdom
from the 1930s to the 1990s, 275
A14-2 The world in 1897, with British possessions marked in red, 282
A17-1 Summary of observations that show the Earth is warming (red arrows)
while the Sun has been constant over the same period of time, 326
A17-2 A comparison of the existing four global surface temperature datasets
that are used in climate analyses, 327
A17-3 Sea-level rise based on radar altimeters from TOPEX and Jason, with
seasonal variations removed, 329
A17-4 A comparison between the total solar irradiance and the NASA/GISS
surface temperature data, both from 1979 to 2010, 330
A17-5 Representation of a general circulation model, 331
A17-6 Change in precipitation between the 1971–2000 average and the
2091–2100 average in inches of liquid water/year, 331
A17-7 Rift Valley fever major outbreak events plotted against time and
the Southern Oscillation Index, a measure of the phase of El Niño/
Southern Oscillation events, 333
A17-8 Summary Rift Valley fever (RVF) risk maps for (A) Eastern Africa:
September 2006–May 2007; (B) Sudan: May 2007–December
2007; (C) Southern Africa: September 2007–May 2008; and (D)
Madagascar: September 2007–May 2008, 335
A17-9 Stem rust symptoms on wheat, 336
A17-10 False-color Landsat satellite data (RGB 642) showing glaciers as the
blue colors. The green colors represent green vegetation and the red
colors represent areas of rock, sand, and soil, 337
xxii TABLES, FIGURES, AND BOXES
A20-1 (A) Greater mouse-eared bat (Myotis myotis) with white fungal growth
around its muzzle, ears, and wing membranes. (B) Scanning electron
micrograph of a bat hair colonized by Geomyces destructans, 371
A20-2 Locations in Europe of bats positive for Geomyces destructans by
PCR alone (circles) or by PCR and culture (solid stars) and bats
negative for G. destructans but positive for other fungi (square), 375
BOXES
WO-1 The Fungal Gardens of Leafcutter Ants, 10
WO-2 Factors in the Emergence of Infectious Diseases, 17
Workshop Overview
1
2 FUNGAL DISEASES
Fungal diseases of plants, animals, and humans have altered tree population
diversity and forest ecosystem dynamics, devastated agricultural crops, triggered
global population declines and extinctions in wildlife, and contributed to death
and disability in humans. Cryptococcus gattii (C. gattii), a pathogenic fungus
that emerged in 1999 on Vancouver Island, British Columbia, Canada, is caus-
ing a growing epidemic of human and animal infections and deaths (Galanis
and MacDougall, 2010). Since its initial recognition, the pathogen has spread
from Vancouver Island to mainland British Columbia and south into the Pacific
Northwest of the United States. This fungal pathogen has been associated with
338 confirmed human infections and 40 deaths1 in these regions, which represents
the largest documented population of C. gattii infected people in the world (Datta
et al., 2009a; Galanis and MacDougall, 2010). Bat white-nose syndrome (WNS)
and amphibian chytridiomycosis2 have caused massive population declines and
threaten local extinctions of New World bat and amphibian species, respectively
(Frick et al., 2010; Skerratt et al., 2007). By 2009, the geographic range of two
virulent and highly aggressive strains3 of yellow “stripe” rust—first detected in
North America in 2000—expanded to include major wheat-producing areas on
five continents, threatening the global wheat supply (Hovmøller et al., 2010).
The recent observation that a fungus (Nosema spp.), in combination with a DNA
virus, might be associated with “colony collapse” disorder—a disease that has
destroyed 20–40 percent of the honeybee colonies in the United States since
2006—underscores the direct and indirect impacts and ecosystem dynamics of
fungal diseases in human, plant, and animal communities (Bromenshenk et al.,
2010).
Fungal organisms interact with humans, animals, and plants in beneficial as
well as pathogenic ways. A dozen fungal diseases are considered “life threaten-
ing” to humans. At the same time, human health has benefited immensely from
fungal-derived antibiotics, such as penicillin (Blackwell et al., 2009; Buckley,
2008; Casadevall, 2007). Indeed, fungi are indispensible to life on this planet
through their ability to break down complex organic matter and recycle essential
nutrients back into the environment (Wainwright, 1992).
The fungal kingdom is among the most diverse kingdoms in the Tree of Life
(Blackwell, 2011). Yet, fewer than 10 percent of fungal organisms have been
formally described (Hawksworth, 1991, 2001). For the purposes of this chapter,
the terms fungi, fungal, and fungus are used inclusively to describe all organisms
traditionally studied by mycologists—including species that are now excluded
from Kingdom Fungi (e.g., Phytophthora spp. which are members of Oomycota)
or whose relationship to the fungal kingdom have yet to be determined (e.g., the
1╛╛As
of December 2010.
2╛╛In
this chapter, we will refer to this disease as amphibian chytridiomycosis and to the associated
pathogen (Batrachochytrium dendrobatidis) as Bd.
3╛╛Puccinia striiformis Westend. f.sp. tritici Eriksson.
WORKSHOP OVERVIEW 3
microsporidia Nosema spp. and the newly discovered cryptomycota) (see Jones
et al., 2011; Stajich et al., 2009).
Despite the extensive influence of fungi on economic well-being, as well as
on human, animal, plant, and ecosystem health, the threats posed by emerging
fungal pathogens are often unappreciated and poorly understood. On December
14 and 15, 2010, the Institute of Medicine’s (IOM’s) Forum on Microbial Threats
hosted a public workshop on this topic in order to explore the scientific and policy
dimensions associated with the causes and consequences of emerging fungal
diseases. Through invited presentations and discussions, the workshop explored
the environmental, host (plant, animal, and human), and pathogen-related factors
influencing the emergence, establishment, and spread of fungal pathogens, as well
as the impacts of these diseases on human and animal health, agriculture, and
biodiversity. Workshop participants also considered and discussed opportunities
to improve surveillance, detection, and response strategies for identifying and
mitigating the impacts of these diseases in order to better prepare for future out-
breaks. Convened in response to the perceived threat posed by emerging fungal
diseases to human, animal, and plant health, this was the first workshop in the
Forum’s 15-year history that focused exclusively on fungal pathogens.
Fungal Diversity
Existing as single-celled organisms, such as yeasts, or complex communities
of filamentous mycelial networks covering hundreds of acres, fungi are ubiqui-
tous in nature and display a dazzling array of sizes, shapes, and colors, including
many that are bioluminescent (Figure WO-2) (Blackwell, 2011; Desjardin et al.,
2010; Lutzoni et al., 2004).
The fungal life cycle is equally varied. Fungi can reproduce asexually or
sexually through life cycles that range from simple to complex—including “di-
morphic” switching between yeast and filamentous forms and the use of multiple
host species (Blackwell et al., 2009). Spores5 are produced during the fungal life
cycle and may be passively or actively dispersed through a variety of environ-
mental media including air, water, wind, animals, and materials (Blackwell et al.,
2009). Fungal growth, reproduction, spore production, and dispersal are also ex-
quisitely sensitive to environmental conditions including temperature, humidity,
4╛╛This number is considered by many to be an underestimate of the actual number of fungal species;
freezing or drying (better than mycelia and yeast cells), for months and even years.
WORKSHOP OVERVIEW 5
Animals (outgroup)
(1 293 642 species)
? Microsporidia
(1300 species)
Chytridiomycota
“ Chytrids”
(706 species)
Neocallimastigomycota
(20 species)
Blastocladiomycota
Fungi
Zygomycota 1
“ Zygomycetes”
? (327 species)
Zygomycota 2
(744 species)
Entomophthorales
? (277 species)
Glomeromycota
(169 species)
Ascomycota
(64 163 species)
Basidiomycota
(31 515 species)
FIGURE WO-1╇ The fungal kingdom. The classification of species within kingdom Fungi
continues to evolve. The diagram above provides an overview of some of the primary
lineages of fungal organisms andFigure WO-1.eps
the estimated number of species for each lineage.
SOURCE: Blackwell (2010).
winds, and water (Bahn et al., 2007; Judelson and Blanco, 2005; Kauserud et al.,
2008; Kumamoto, 2008).
Fungi are highly adaptable to new environmental niches including what
might be considered “extreme” environments (Gostinčar et al., 2010; Le Calvez
et al., 2009). Some have suggested the ability of fungi to access multiple strate-
gies for reproduction contributes to why fungi are so “adept at adaptation.” Under
different environmental conditions, fungal reproduction can maintain character-
istics adapted to a particular environmental niche or generate genetically diverse
offspring that can quickly respond to changing host or environmental factors
(Heitman, 2006). (Dr. Blackwell’s contribution to the workshop summary report
can be found in Appendix A, pages 116–167.) Keynote speaker Meredith Black-
well, of Louisiana State University, noted that scientists continue to find new
species of fungi in a wide range of environments—from tropical and temperate
forests to the guts of insects (e.g., Arnold et al., 2003; Gostinčar et al., 2010;
Miller et al., 2001; Suh and Blackwell, 2006). These discoveries often reveal the
unique capabilities of these microorganisms. As observed by Casadevall, some
fungal species can survive and thrive in high radiation and other extreme envi-
ronments. Zhdanova et al. (2000) reported extensive fungal growth on the walls
6 FUNGAL DISEASES
A B C
D E F
G H
I
.
WO-2 new
WORKSHOP OVERVIEW 7
FIGURE WO-2╇ Diversity of fungal morphology. (A) Two flagellated fungal cells from
the recently discovered group of fungi known as cryptomycota. This ancient group of
organisms is thought to be distinct from other fungi because of the absence of a cell wall
made of chitin; (B) asexual, spore-producing culture of Cryphonectria parasitica (chestnut
blight fungus); (C–F) multicellular, spore-producing structures (fruiting bodies) are pro-
duced during the sexual phase of the fungal life cycle. Many fruiting bodies are familiar
as mushrooms—including species that are consumed by humans as food: (C) Morchella
conica (morel) and (D) Crucibulum laeve (bird’s nest fungus). Mushrooms of some species
are known to be toxic or poisonous to humans: (E) Amanita muscaria. Fungal fruiting bod-
ies can exhibit a wide range of shapes and sizes, including (F) the bioluminescent “shelf”
fungus, Panellus stipticus; (G) Micrograph of Phytophthora ramorum chlamydospores;
(H) SEM photomicrograph prepared from G. destructans culture isolated from bat tissue
samples collected from Williams Hotel Mine; note curved conidia borne in whorls on
septate hyphae; bar is 2 µm. All images are pseudo-colored in Adobe Photoshop 9.0; (I)
“fairy rings” in which mushrooms sprout along the outer edge of a sprawling, underground
mycelial network. These networks (mycelia) have been known to cover several hundred
acres. One of largest known mycelia has been estimated to encircle 900 hectares (3.4
square miles).
SOURCE: (A) Micrograph kindly provided by Meredith Jones, Exeter University; (B)
photo by Kent Loeffler, provided by Alice C.L. Churchill, Cornell University; (C–F) Wi-
kimedia Commons; (G) photo provided courtesy of Paul Reeser, Oregon State University;
(H) Chaturvedi et al. (2010); (I) Wikimedia Commons.
and other areas of the shelter installed around the damaged unit of the Chernobyl
nuclear power plant, including 37 species among 19 genera 6; fungi are also
known to inhabit high-radiation space environments and have even colonized the
International Space Station (Dadachova and Casadevall, 2008).
The fungal pathogen responsible for sudden oak death and ramorum blight,
Phytophthora ramorum, was only identified as a new species in 2000. Since
then, according to speaker David Rizzo of the University of California at Davis,
researchers have identified an additional 50 Phytophthora7 species. (Dr. Rizzo’s
contribution to the workshop summary report can be found in Appendix A, pages
312–324.) As Rizzo observed, these new discoveries do not reflect recent fungal
evolution, but are a reflection of the fact that “we just haven’t really been looking
6╛╛Many of the species inhabiting the most heavily contaminated sites of the Chernobyl nuclear
power plant were rich in melanin (a high molecular weight pigment). Dadachova et al. (2007) reported
that radiation enhances the growth of melanized Wangiella dermatitidis, Cryptococcus neoformans,
and Cladosporium sphaerospermum cells.
7╛╛Phytophthora (“plant destroyer”) is a genus of approximately 100 species that includes several
notorious plant pathogens, including Phytophthora infestans, which caused the Irish Potato Famine.
Phytophthora species are oomycetes, which are fungus-like organisms in the kingdom Stramenopila.
8 FUNGAL DISEASES
for them.” Several other forest fungi that have caused major damage in the past,
including the fungi responsible for chestnut blight and Dutch elm disease, were
unknown to science until they started causing noticeable damage and die-off of
forest and urban trees (Brasier and Webber, 2010).
8╛╛Services
provided by ecosystems that benefit humans and are necessary for a healthy planet like
oxygen production, water purification, pollination, soil formation, and nutrient recycling. See www.
conservation.org/resources/glossary/Pages/e.aspx (accessed on June 13, 2011).
9╛╛Deriving nutrients from dead organic matter.
10╛╛Not dependent on a host for survival.
11╛╛For example: A parasitic fungus, Ophiocordyceps sinensis, grows in the Tibetan Plateau in China
and is highly valued for its “purported medicinal benefits,” including uses as “a treatment for cancer
and aging and as a libido booster.” The nutty-tasting fungus is considered “fungal gold” because it
can be sold for high prices in Chinese markets (see Roach, 2011).
12╛╛Other medicines such as the immunosuppressant cyclosporine A and statin drugs also are derived
from fungi.
WORKSHOP OVERVIEW 9
nutrition of many insects (e.g., Nardi et al., 2006; Suh et al., 2003, 2005). Fungi
also are cultivated by fungus-farming termites and ants (Aanen et al., 2002;
Currie et al., 2003; Dentinger et al., 2009; Munkacsi et al., 2004) (Box WO-1).
Not all fungal–insect associations are mutualistic. Blackwell described the
parasitic but not usually pathogenic fungi in the order Laboulbeniales. She noted
the reports of extreme host specificity exhibited by different species in this
order—sometimes inhabiting only certain parts of the host insect (Weir and
Beakes, 1995). Most laboulbenialean species are associated with beetles (Cole-
optera), and flies (Diptera), but they are also associated with a diverse array of
host species in other insect orders, mites and millipedes (Weir and Beakes, 1995).
Blackwell discussed a number of fungal–plant symbioses. She estimated
that:
Endophytes—fungi that live inside the plant tissue but without causing
any obvious negative effects—are less well known than other plant–fungal as-
sociations, but mycologists find them wherever they look (Arnold et al., 2003;
Rodriguez et al., 2009). Numerous endophytic fungal infections have been ob-
served in cocoa trees (Theobroma cacao) and they may play an important role
in host defense by decreasing the damage associated with Phytophthora spp.
infections (Arnold et al., 2003). To illustrate the complexity of these relationships,
Blackwell noted interactions among the fungus Curvularia protuberata, the grass
Dichanthelium lanuginosum,13 and a fungal virus. The grass infected with the
fungus infected with “Curvularia thermal tolerance virus” provides thermal re-
sistance benefits for the host plant. This tripartite relationship allows the grass to
grow in the high-temperature soils of Yellowstone National Park (Márquez et al.,
2007). Blackwell pointed to the red-cockaded woodpecker (Picoides borealis)
as just one example of the many ways that fungi confer benefits to the health of
ecosystems. These woodpeckers usually nest in trees infected with red heart rot
(Phellinus pini) (Hooper et al., 1991).
BOX WO-1
The Fungal Gardens of Leafcutter Ants
Over the past 50 million years, a unique symbiosis has developed between
attine (fungal growing) “leafcutter” ants and fungi in the Lepiotacea family. In what
biologists consider the earliest form of agriculture, leafcutter ant colonies grow
and meticulously maintain a specific fungal cultivar for food (Schultz and Brady,
2008; Wade, 1999).
Inhabiting forest ecosystems throughout Mexico and Central and South Amer-
ica, these ant colonies can number more than 8 million individuals. Foraging ants
bring cut pieces of leaves back to the colony where they are broken down and fed
to the fungus by worker ants (see Figure WO-1-1).
A second symbiotic relationship protects these fungal gardens. Pseudono-
cardia bacteria, which grow on the bodies of the worker ants, produce antibiotic
compounds that prevent the growth of parasitic molds (Currie et al., 1999).
Figure WO-QRtode.eps
bitmap
WORKSHOP OVERVIEW 11
Fungi as Pathogens
The longstanding utility of fungi to all life on earth has often been matched
by their ability to directly or indirectly cause devastating disease in human,
animal, and plant hosts. Fungi are the predominant pathogen species in plants,
remarked Casadevall, and fungi can also cause disease in healthy humans and
animals. Described by several workshop participants as “formidable pathogens,”
many fungi can also endure adverse environmental conditions and thrive outside
of their host (Casadevall, 2007).
Fungal pathogens in general execute a series of sequential steps in order to
cause disease, remarked speaker Barbara Howlett of the University of Melbourne.
(Dr. Howlett’s contribution to the workshop summary report can be found in Ap-
pendix A, pages 264–273.) These pathogens must:
Very few fungal pathogens are able to cause disease in hosts from the plant
and animal kingdoms; those that do are referred to as trans-kingdom pathogens
(De Lucca, 2007).15 Fungi can also form different associations with different host
types. For example, the fungus Cryptococcus gattii is pathogenic in animals in-
cluding humans, but forms non-pathogenic associations with plants –which play
an essential role in the maintenance of C. gattii spores in certain environmental
niches (Bartlett et al., 2007; Xue et al., 2007).Once outside of a host, fungal
pathogens of animals and plants often have different requirements for survival.
Animal pathogens, noted Howlett, are often soil saprophytes that are free-living
rather than obligate.16 In contrast, some plant pathogens can only survive on the
tissue of a specific plant host(s).
14╛╛For more information, see contributed manuscript by Barbara Howlett in Appendix A (pages
264–273).
15╛╛Howlett noted two trans-kingdom pathogens during her remarks: Fusarium oxysporum f. sp.
lycopersici, which causes vascular wilt in plants and is an emerging human pathogen (Ortoneda et al.,
2004); and Aspergillus flavus, which infects corn and is an emerging pathogen in immunocompro-
mised humans (Krishnan et al., 2009).
16╛╛Capable of existing only in a particular environment; an obligate parasite cannot survive inde-
Fungal plant diseases have far-reaching health impacts that extend beyond
the infected plant species—including, but not limited to, negative impacts on
associated flora and fauna (Giraud et al., 2010; Loo, 2009). As the Irish Potato
Famine illustrated, crop losses can have devastating impacts on populations that
are heavily, or solely, dependent on a single food source for their caloric needs.
Speaker Jim Stack of Kansas State University observed that with 59 percent of
calories consumed by humans being derived from just four plant species (rice,
wheat, maize, and potatoes), fungal diseases in these staple crops may cata-
strophically threaten local and global food security (Strange and Scott, 2005;
Vurro et al., 2010). (Dr. Stack’s contribution to the workshop summary report
can be found in Appendix A, pages 273–296.)
17╛╛Crop losses due to all pathogens (1988–1990) totaled $33 billion for rice, $14 billion for wheat,
$7.8 billion for maize, and $9.8 billion for potatoes (Oerke et al., 1995; Rosenzweig et al., 2001).
WORKSHOP OVERVIEW 13
FIGURE WO-3╇ Depiction of starving Irish children in 1847 potato famine; by Cork artist
James Mahony (1810–1879). Figure WO-3.eps
SOURCE: Wikimedia Commons.
bitmap
14 FUNGAL DISEASES
of humans have had a lower disease burden than bacterial, viral, or parasitic
infections, although this disease burden may be changing. It has been noted that
fungal diseases are increasing in incidence in the growing populations of immu-
nocompromised human hosts (Romani, 2004). Once established, fungal diseases
are often difficult to treat (Casadevall, 2007; Romani, 2004).
Disease in humans most often results from opportunistic18 infections (Shoham
and Levitz, 2005). Only a few fungal diseases (e.g., coccidioidomycosis, histo-
plasmosis) are caused by “primary” fungal pathogens19 that induce symptomatic
disease in otherwise healthy people (Casadevall, 2007; Cutler et al., 2007). The
“apparent” resistance of humans to fungal disease may be a reflection of the host
immune response, coupled with the high basal temperature of mammals, which
often exceeds the thermotolerance20 range for many fungi (Casadevall, 2005;
Garcia-Solache and Casadevall, 2010; Robert and Casadevall, 2009).
Primary fungal pathogens of humans can also infect other mammals, such as
domesticated livestock and companion animals. These diseases are generally not
considered contagious and are acquired via inhalation of aerosolized infectious
propagules21 from environmental reservoirs, such as soil or trees (Casadevall and
Pirofski, 2007). According to speaker Luis Padilla of the Smithsonian Conserva-
tion Biology Institute, wildlife are also affected by opportunistic and primary
fungal pathogens, but the epidemiology of these diseases in wildlife is not well
understood. (Dr. Padilla’s contribution to the workshop summary report can be
found in Appendix A, pages 296–312.) Two fungal diseases of wildlife, am-
phibian chytridiomycosis and bat white-nose syndrome, emerged rapidly and
unexpectedly over the past several decades. These diseases are associated with
unprecedented local and global population declines of amphibian and bat species,
and pose serious threats to biodiversity and ecosystem stability (Frick et al., 2010;
Wake and Vredenburg, 2008).
18╛╛Resulting from pathogen entry via wounds or weakened state of the host, or as a disturbance of
portunists rarely cause disease in an immunocompetent host whereas the primary pathogens do. For
more information see: Cutler et al. (2007).
20╛╛Garcia-Solache and Casadevall (2010) define thermotolerance as the ability to grow at mam-
malian (37°C) and higher temperatures. Most fungi thrive in the range of 12°C to 30°C, but there are
wide temperature tolerances among species, with some growing at temperatures as low as –10°C or
as high as 65°C. See contributed manuscript by Casadevall in Appendix A (pages 181–188).
21╛╛Spores or encapsulated yeast cells.
WORKSHOP OVERVIEW 15
due, in part, to the fact that “when [fungal pathogens] get into an ecosystem—a
vertebrate host, for example—they simply don’t care. They have no need for that
host in order to go forward. They will take down every last member of the spe-
cies.” In contrast, most newly introduced viral and bacterial pathogens in a naïve
host eventually attenuate their virulence such that infection does not kill the host.
Such adaptations are beneficial to both the host and pathogen in that the host sur-
vives and the pathogen avoids an “evolutionary dead end” (IOM, 2009). As noted
by Casadevall and Pirofski (2007), the host independence of “environmental”
microbes,22 including many fungi, may confer advantages that promote survival
and virulence in other niches, including new ecosystems and novel host species.
The term “invasive species” is used to describe “non-native”23 plants and
animals that, when introduced to new environments, reproduce or spread so ag-
gressively that they harm their adopted ecosystems (Carlton, 2004; Dybas, 2004).
They compete with native organisms for food and habitat, act as predators or
parasites of native species, and cause or carry diseases, often with devastating
ecological and economic consequences (Pimentel et al., 2005). As observed by
Morse (2004), infectious diseases represent another form of biological invasion—
often arising “out of nowhere” with devastating effects.
Discussions during the workshop illuminated the capacity of many fungal
pathogens to persist in environmental reservoirs and to readily adapt to new envi-
ronmental niches and host species. Like invasive species, these fungal pathogens
have been able to thrive in new environments and are changing the ecosystem in
ways that are difficult to anticipate and even more daunting to prevent (Desprez-
Loustau et al., 2007; Giraud et al., 2010; Rizzo, 2005). Given both the links and
similarities between invasive species and many pathogenic fungi, it may be useful
to view the origins of disease emergence, and the strategies deployed to prevent
or mitigate the threats associated with fungal pathogens, through the larger lens
of biological invasiveness.
FACTORS OF EMERGENCE
Diseases are categorized as “emerging” if their incidence24 or virulence25 has
recently increased or if they begin to infect a novel host or population (WHO,
2010). As illustrated in Figure WO-4, disease26 results from a complex interplay
of interactions among the pathogen, host, and environment.
22╛╛Microbes acquired from the environment (in contrast to acquisition from other living hosts)
(Casadevall and Pirofski, 2007).
23╛╛Also called “exotic,” “alien,” and “nonindigenous” species.
24╛╛As used in epidemiology, the number of new cases of a disease that occur in a defined population
infection has become clinically apparent (IOM, 1992). Some exposures to infectious disease-causing
agents can also produce asymptomatic illnesses that can be spread to others.
16 FUNGAL DISEASES
FIGURE WO-4╇ The epidemiological triad. The familiar “epidemiological triad” concept
(host–pathogen–environment), as illustrated in the famous diagram of Snieszko (1974),
FigureofWO-4.eps
neatly illustrates the complex interplay factors that result in disease at the individual
and population levels. The presence of abitmap
pathogen is a necessary, but not sufficient, cause
of a particular disease (IOM, 2008b).
SOURCE: Snieszko (1974, Figure 1).
BOX WO-2
Factors in the Emergence of Infectious Diseases
pathogens27 (Brasier, 2000). As Rizzo noted, “for some of these tree pathogens,
I don’t think there is anything extraordinary about the pathogens themselves.”
Rather, “it is the movement of pathogens from one environment to another that
seems to be driving much of the destruction.”
As described below, discussion at the workshop considered the influence of
human activity and behavior, winds and weather, host susceptibility, and patho-
gen adaptation and change on fungal disease emergence.
27╛╛Stack
explained that interspecific hybridization is another unexpected outcome of pathogen glo-
balization. Two different (and previously isolated) fungal species mate and produce novel “hybrid”
offspring. Fungal pathogens of plants produced by interspecific hybridization are more aggressive
than either parental phenotype and may occupy a new host range (see Brasier, 2000). Stack noted
that even in a single nursery, the normal process of taking care of plants, which includes watering,
can result in a water splash that brings two different fungal species together in a single pot, where
interspecific hybridization can occur.
18 FUNGAL DISEASES
hosts to infectious disease agents (Anderson et al., 2004; Brasier, 2008; Daszak
et al., 2000). Travel, trade, migration, agricultural practices, and land use patterns
have all contributed to increased opportunities for contact between introduced
pathogens and naïve and susceptible host populations (IOM, 2010, and refer-
ences within).
Speaker Matthew Fisher of the Imperial College London, remarked that
migrating humans have been globalizing pathogens for thousands of years. (Dr.
Fisher’s contribution to the workshop summary report can be found in Appendix
A, pages 355–367.) He pointed to the spread of Coccidioides posadasii that ac-
companied human migration between 5,000 and 10,000 years ago through North,
Central, and South America as an example (Fisher et al., 2001). Stack agreed,
noting that “global trade is not new; we have had 3,000 years of global trade.” He
went on to state that, “what is new is the magnitude of trade in plants and plant
products and the speed at which they move around the world” [emphasis added].
28╛╛According to Stack (2010), each year, 12,000–14,000 potential pathogen and pest problems are
more than 100 plant species (Grünwald et al., 2008). Stack noted that potential
pathogens can be transported in plants, plant associated material (e.g., soil),
seeds, and objects manufactured using plant products, such as wooden instru-
ments and packing materials.
Human spatial mobility has increased at least 1,000-fold in the past 200
years, with more people traveling faster, farther, and less expensively than ever
(Figure WO-5) (Cliff and Haggett, 2004; Hufnagel et al., 2004).
Travelers are now able to easily explore once-remote areas that serve as
both sources and sinks for emerging infectious diseases (Choffnes, 2008; IOM,
2010). Adventure travelers intrude on once-remote environments and often make
contact with exotic wildlife, encountering microbes that have never before been
recognized as human pathogens in the “developed” world (IOM, 2010). These
ecotourists become unwitting vectors of disease when they bring these exotic in-
fectious diseases back with them—on their person/clothing/luggage, etc.—when
they return to their home countries. If the conditions are favorable, an introduced
pathogen may persist and spread (Wilson, 2003). White-nose syndrome, which
is currently decimating New World bat populations in the United States, may
have been accidentally introduced by recreational cavers from Europe29 (Wibbelt
et al., 2010).
Infectious disease pandemics have also been associated with the legal and
illegal trade in and transportation of animals (IOM, 2010; Karesh et al., 2005;
Smith et al., 2009). Between 2000 and 2006, the United States traded approxi-
mately 1.5 billion animals, according to speaker and Forum member Peter Daszak
of EcoHealth Alliance. (Dr. Daszak’s contribution to the workshop summary
report can be found in Appendix A, pages 188-196.) These animals come from
a wide range of species, and most animal imports into the United States come
from emerging infectious disease “hot spots” (see Table WO-1) (Jones et al.,
2008; Smith et al., 2009).
The international amphibian trade is thought to have contributed to the emer-
gence and global spread of amphibian chytridiomycosis (Catenazzi et al., 2010;
Daszak et al., 2003; Fisher and Garner, 2007; Schloegel et al., 2010; Weldon
et al., 2004). Since the 1990s, this fungal disease has been implicated in the wide-
spread population declines—including some local extinction events—of more
than 200 species of frogs, toads, and salamanders (Fisher et al., 2009; Kilpatrick
et al., 2009; Lips et al., 2006; Schloegel et al., 2006; Skerratt et al., 2007).
Speaker Ché Weldon of North-West University of South Africa noted the
many points in the global amphibian trade pathway where traded species from
different origins come into contact including collector-supplier facilities, breed-
ing facilities, end-user facilities, etc. (Dr. Weldon’s contribution to the workshop
summary report can be found in Appendix A, pages 355–367.) Weldon discussed
two widely traded amphibian species—the African clawed frog, Xenopus laevis,
29╛╛Other possible explanations include the importation of horticultural soils from Europe.
20
FIGURE WO-5╇ Global aviation network. A geographical representation of the civil aviation traffic among the 500 largest international air-
ports in 100 countries is shown. Each line represents a direct connection between airports. The color reflects the number of passengers per day
traveling between two airports, with the most intense traffic (25,000) noted in yellow.
SOURCE: Hufnagel et al. (2004).
WORKSHOP OVERVIEW 21
30╛╛Includes
weather phenomena that are at the extremes of the historical distribution, especially
severe or unseasonable weather (e.g., extreme heat or cold, tropical cyclones, tornadoes). http://
en.wikipedia.org/wiki/Extreme_weather.
22 FUNGAL DISEASES
FIGURE WO-6╇Selected dispersal events of fungal pathogens. Red and blue arrows
Figure
indicate invasions of new territories WO-6.eps
(first year recorded in brackets). Red arrows indicate
dispersal that probably occurred by direct movements of airborne spores (I, II, III, and
bitmap
IV). Blue arrows indicate pathogens that were probably transported to the new territory
in infected plant material or by people and spread thereafter as airborne spores (V, VI,
VII, and VIII). Orange circles indicate the worldwide spread of black Sigatoka disease of
banana; the first outbreak on each continent is marked (IX). Green arrows indicate periodic
migrations of airborne spores in extinction-recolonization cycles (X, XI, XII, XIII, XIV).
SOURCE: From J. K. M. Brown and M. S. Hovmøller. 2002. Aerial dispersal of pathogens
on the global and continental scales and its impact on plant disease. Science 29(5581):537–
541, reprinted with permission from AAAS. Background map provided courtesy of
Christopher Lukinbeal, University of Arizona.
gens are soil-associated, wind and other factors associated with soil disturbances
can disperse spore-associated dusts into the air. Once airborne, spores may pas-
sively travel on the wind over great distances—often hundreds or thousands of
miles—to new geographic areas and new host environments (Figure WO-6 [red])
(Brown and Hovmøller, 2002).
et al., 2003; Crum-Canflone, 2007; Warnock, 2006). In 1994, the magnitude 6.7
Northridge earthquake led to an outbreak of valley fever in southern California
(Figure WO-7A) (Schneider et al., 1997). More recently, the massive dust storm
that swept through Arizona on July 5, 2011 is predicted to cause a similar increase
in cases of valley fever (Chan, 2011) (Figure WO-7B). Speaker John Galgiani,31
of the University of Arizona, explained that even small winds or soil disturbances
can easily loft spore-laden dusts into the air. (Dr. Galgiani’s contribution to the
workshop summary report can be found in Appendix A, pages 196–207.) Galgiani
further observed that inhalation of a single spore “at the right time” can cause
disease. Approximately 40 percent of infected persons develop symptoms, which
initially manifest as pneumonia (i.e., cough, chest pain, fever, and weight loss);
fatigue; bone and joint pains (“desert rheumatism”); or skin rashes (Hector and
Laniado-Laborin, 2005; Tsang et al., 2010). While dust storms and environmental
disturbances are clearly an important driver of the spread of Coccidioides spp.,
Galgiani said that simply living in an endemic region,32 without any direct contact
with the soil, puts one at risk of exposure. Yet the fungus is only sparsely distrib-
uted. Galgiani noted that, “you can do a lot of desert digging and disrupting and
not even be close to the fungus.”33
Airborne spore dispersal may also synergize with intercontinental trade and
travel to rapidly spread diseases between and within continents (Figure WO-6
[blue]). Yellow rust (Puccinia striiformis f. sp. tritici) is believed to have been
introduced into Western Australia from southern Europe, in 1979, as an adherent
spore on an air traveler’s clothing (Wellings, 2007). Once introduced into Aus-
tralia, the pathogen spread across Australia’s wheat belt and into New Zealand
via wind dispersal (Brown and Hovmøller, 2002). Indeed, winds allow many ag-
riculturally important fungal plant diseases to gradually expand their geographic
range (Brown and Hovmøller, 2002).
Pandemics caused by intercontinental aerial dispersal of spores can and do
occur—often facilitated by hurricanes and other extreme weather events. Ex-
amples include:
31╛╛Dr. Galgiani is also Chief Medical Officer at Valley Fever Solutions, Inc. which has licensed the
development of nikkomycin Z as a treatment for valley fever from the University of Arizona.
32╛╛Although widely perceived as endemic to the southwestern United States, Galgiani observed that
the endemicity of the disease extends through Mexico into Central and some parts of South America
(Tsang et al., 2010).
33╛╛It has been suggested that this spotty distribution is a result of the abundant fungal sporulation
that may accompany fungal decomposition of infected animal remains (whether or not fungal infec-
tion was responsible for an animal’s death) (Sharpton et al., 2009).
24 FUNGAL DISEASES
Figure WO-7.eps
B bitmap
FIGURE WO-7 Environmental disturbances and dust storms contribute to the dispersal
of fungal spores. (A) Dust from landslides caused by a 5.6 magnitude aftershock of the
1994 Northridge earthquake blows out of the Santa Susana Mountains into the Simi Valley.
An outbreak of valley fever occurred in the Simi Valley following the January 17, 1994,
6.7 magnitude Northridge earthquake. (B) The leading edge of a violent dust and sand
storm (known as a haboob). The wall of dust and sand that swept through Arizona on July
5, 2011, was estimated to be more than 50 miles wide. The storm travelled over 150 miles
and reached peak heights of 8,000 to 10,000 feet.
SOURCE: Photos courtesy of Tom Freeman; National Oceanic and Atmospheric
Administration.
WORKSHOP OVERVIEW 25
• Coffee leaf rust (Hemileia vastatrix) may have been transported via trans-
atlantic winds between Angola to Bahia in Brazil in 1970 (Figure WO-6
[red]) (Brown and Hovmøller, 2002).
• Asian soybean rust was brought into the United States from South Amer-
ica by Hurricane Ivan in 2004 (Schneider et al., 2005).
Some scientists are concerned that the frequency and duration of such ex-
treme weather events could increase with global climate change, which in turn
could influence the incidence and intensity of fungal disease outbreaks (Garrett
et al., 2006; Greer et al., 2008).
or “water mold” that belongs to the Kingdom Stramenopila (a major eukaryotic group that includes
diatoms and brown algae, and is distinct from plants, fungi, and animals). Like fungi, oomycetes
“exhibit filamentous growth, produce sexual and asexual spores, and can feed on decaying matter or
be obligate parasites of plants” (Kliejunas, 2010).
26 FUNGAL DISEASES
experiencing total crop loss. Average yield loss in the Corn Belt states 36 was
20–30 percent, with some parts of Illinois and Indiana reporting yield losses of
50–100 percent (Ullstrup, 1972). In the 1970 season alone, the SCLB epidemic
led to the loss of 710 million bushels of corn—valued at more than $1 billion at
the time (or about $5.6 billion in 2009 dollars) (Tatum, 1971).
Compton Tucker, of the National Aeronautics and Space Administration
(NASA) Goddard Space Flight Center, presented data from a variety of satellite
and ground sources37 documenting increases in global temperatures worldwide,
as well as changes in the atmospheric concentration of carbon dioxide. (Dr.
Tucker’s contribution to the workshop summary report can be found in Appendix
A, pages 324-342.) He also explained how general circulation models, which
simulate the atmosphere, accounting for wind, humidity, clouds, temperature,
composition of the atmosphere (e.g., presence of trace gases), and other weather-
related variables, can be used to predict where on the surface of the earth (both
land and water) temperature and precipitation levels are likely to change.38 Ac-
cording to Tucker, these models predict that over the next century, average surface
temperatures will increase by 2–5°C, and regions of the world will get wetter or
drier (Figure WO-8).
Fungal diseases are influenced by weather fluctuations and display
“seasonality”—suggesting the possible influence of long-term climate changes
(IOM, 2003, 2008a; Rosenzweig et al., 2001). Stack noted that the onset of potato
late blight has been occurring earlier and earlier over the past 20 years in some
regions of the world and has resulted in more severe losses and greater mitigation
challenges (Hannukkala et al., 2007). In part, this is due to changing temperatures
and increased frequency of precipitation (Hannukkala et al., 2007).
Stack observed that modeling studies predict many negative impacts on plant
health in response to climate change, including shifts in the range, timing, and
severity of fungal diseases of plants39 (Jeger and Pautasso, 2008; Pautasso et al.,
2010). A 3°C increase in temperature, for example, is anticipated to alter the
phenology40 and conditions of the host species enough to result in expansion of
the geographic range of Phytophthora cinnamomi, which has already decimated
forests across southeastern Australia (Lonsdale and Gibbs, 1996). An enormous
effect is predicted for the severity of phoma stem canker (Leptosphaeria macu-
36╛╛The area in the Midwestern United States, roughly covering western Indiana, Illinois, Iowa, Mis-
souri, eastern Nebraska, and Eastern Kansas, in which corn (maize) and soybeans are the predominant
field crops (Encyclopedia Britannica: eb.com).
37╛╛These data include NASA satellite data on solar irradiance (i.e., energy output of the sun); Na-
tional Oceanic and Atmospheric Administration, NASA, and other surface data on land and ocean
temperatures worldwide; U.S. military and other satellite and ground data on sea ice; sea-level data;
NASA gravity data; and data on the atmospheric concentration of carbon dioxide and other compo-
nents of the atmosphere.
38╛╛See also contributed manuscript by Tucker in Appendix A (pages 324–342).
39╛╛See contributed manuscript by Jeger in Appendix A (page 273–296).
40╛╛The scientific study of cyclical biological events, such as flowering, breeding, and migration.
WORKSHOP OVERVIEW 27
FIGURE WO-8╇ Change in precipitation between the 1971–2000 average and the 2091–
Figure WO-8.eps
2100 average in inches of liquid water/year.
SOURCE: Geophysical Fluid Dynamicsbitmap Laboratory, National Oceanic and Atmospheric
Administration.
lans) on oilseed rape, with many regions of the United Kingdom expected to
experience a 40–50 percent yield loss by 2050 (Butterworth et al., 2010).
Modeling studies predict that it is not just the plant pathogens themselves
that are likely to be impacted by continued climate change, Stack observed, but
host species as well (Loustau, 2006; Pautasso et al., 2010). Stack remarked that
while modeling studies forecast climate change effects on the distribution or
severity of many fungal plant pathogens, for most crop plants the future is un-
certain—both with regard to plant disease occurrence and the associated impacts
on food security.
41╛╛Immune response (of both vertebrates and invertebrates) to a pathogen that involves the preex-
isting defenses of the body (e.g., barriers formed by skin and mucosa, antimicrobial molecules and
phagocytes). Such a response is not specific for the pathogen (Alberts et al., 2002).
42╛╛Response of the vertebrate immune system to a specific antigen that typically generates immu-
underlie two serious diseases associated with fungal infection: Job’s syndrome 45
and severe coccidioidiomycosis (Buckley et al., 1972; Davis et al., 1996; Holland
et al., 2007; Vinh et al., 2009).
The incidence of opportunistic fungal infections46 has increased recently
and is associated with the growing populations of vulnerable, immunocompro-
mised individuals (e.g., people living with HIV/AIDS, recent organ transplant
recipients) (Romani, 2004). In Casadevall’s opinion, the period since the 1950s
should be viewed as a transition decade in which “fungi become more important
to human health” (Figure WO-9).
In the 1950s, only about 100 reported cases of disease were caused by Cryp-
tococcus neoformans, Casadevall observed; today, there are about 1 million cases
worldwide, mostly among persons with HIV/AIDS (Park et al., 2009). The yeast
infection caused by Candida spp. was also uncommon until the 1950s. Many
have associated the increase in Candida infections to the increased number of
immunocompromised individuals (Dixon et al., 1996). Casadevall speculated
that this may also be linked to the introduction of antibiotics, which altered the
microbial flora in the human host.
Forum member Fred Sparling of the University of North Carolina, Chapel
Hill, remarked “that there was significant cryptococcal disease47 in the pre-HIV
era,” and that he continued to observe the disease in apparently healthy individu-
als. Holland agreed and noted that he expected that “we may find new mecha-
nisms for susceptibility that might not be ‘Mendelian,’48 because it is not familial,
but something that comes on, typically, in adulthood.”
Casadevall considers host immune status in humans so important in the
development of fungal disease that, in his opinion, fungal virulence can only be
properly defined as a function of it. Casadevall went on to explain that pathoge-
nicity is not an invariant, absolute quality in an infectious disease agent, but that
the pathogenicity of a microorganism varies depending on the host and over time
(Casadevall, 2007). He reviewed the “damage response framework”—illustrated
in Figure WO-10—that was developed by Pirofski and Casadevall as a way to
illustrate these concepts (Casadevall and Pirofski, 2003; Pirofski and Casadevall,
2008).
Host damage can derive from either the pathogen (e.g., among immuno-
45╛╛A rare, inherited disease associated with abnormalities of the skin, sinuses, lungs, bones, and
teeth. People with this condition have chronic and severe skin infections (also known as hyper
immunoglobulin E [IgE] syndrome). MedlinePlus: http://www.nlm.nih.gov/medlineplus/ency/
article/001311.htm.
46╛╛These opportunistic fungal diseases include invasive aspergillosis and aspergilloma (Aspergillus
spp.), invasive fusariosis (Fusarium spp.), Pneumocystis pneumonia (Pneumocystis jirovecii), and
invasive candidiasis (Candida sp.) (Nucci and Marr, 2005; Pfaller and Diekema, 2010).
47╛╛Disease caused by Cryptococcus neoformans or gattii infection.
48╛╛A single gene disorder caused by a defect in one particular gene, and characterized by how they
HIV pandemic
Fungal
diseases
VULNERABLE HOSTS
Anti-neoplastic therapy
Anti-inflammatory therapy
Organ transplantation Antibiotic
resistance
Antibiotics
FIGURE WO-9╇ Incidence of systemic fungal disease has increased since the 1950s. Over
the same period, the use of medical technology and the HIV/AIDS pandemic have led to an
Figure WO-9.eps
increased number of immunocompromised individuals. The emergence of systemic fungal
disease in humans is considered by many to be a 20th-century phenomenon.
SOURCE: Adapted from Casadevall (2010).
compromised individuals with weak immune systems) or the host (e.g., among
healthy host individuals whose microbial flora has been disturbed by antibiotic
use, triggering a disproportionately strong immune response) (Casadevall and
Pirofski, 2003). This model predicts that not just immunocompromised individu-
als are at risk of disease from fungal infection(s), but also healthy hosts who
mount a disproportionately strong immune response (Casadevall and Pirofski,
2003).
Casadevall also suggested that there may be additional “subtle” effects of
fungal infection that we are just beginning to observe on a population level.
WORKSHOP OVERVIEW 31
c. 1999 2010
Cryptococcosis HIV-associated cryptococcosis
Transplant-associated cryptococcosis
Sporadic (?)
Host damage
Host damage
Asthma
IRIS-associated cryptococcosis
Disease Disease
Host damage
Host damage
Vaginitis
Oral thrush
Mucocutaneous disease
Disease Disease
He noted that “we are dealing with things now that we never saw 30–40 years
ago. The elimination of many viral and bacterial exposures, especially early in
life, without the concomitant elimination of fungal diseases could be a factor in
asthma and other atopic diseases.” Relman added, “We are not good at measuring
subtle damage. If there are fundamentally important but less obvious forms of
damage going on in the environment due to the emergence of fungi, we are not
going to be very swift at detecting them, or insightful about understanding their
implications.” Infectious propagules49 of C. neoformans spp. are everywhere,
Casadevall stated, and “we are all exposed to them.” Despite this high level of
Microbial Flora
Host immune defenses against fungal disease extend to their microbial flora.
As speaker Vance Vredenburg of San Francisco State University explained, am-
phibians “wear their defenses on their skin” (e.g., glands produce defensive
toxins). (Dr. Vredenburg’s contribution to the workshop summary report can be
found in Appendix A, pages 342–355.) Indeed, Brucker et al. (2008) isolated a
strain of bacteria (Janthinobacterium lividum) from the skin of the red-backed
salamander (Plethodon cinereus) and demonstrated that the bacteria produced
antifungal metabolites at concentrations lethal to the causative agent of amphibian
chytridiomycosis (Bd) (Figure WO-11).
most could not grow at mammalian temperatures, and that every “1°C increase in the 30°C–40°C
range excluded an additional 6 percent of fungal isolates,” implying that fever could significantly
increase the thermal exclusion zone. This led them to conclude that, “Mammalian endothermy and
homeothermy are potent nonspecific defenses against most fungi that could have provided a strong
evolutionary survival advantage against fungal diseases.” See contributed manuscripts by Casadevall
in Appendix A (pages 177–188).
WORKSHOP OVERVIEW 33
Pathogen Adaptation
Many fungi do not need a living host to survive. Fungi are well adapted to
exploit winds and water as a means for their dispersal. Moreover, a variety of
“environmental” cues trigger fungal growth, sexual and asexual reproduction,
sporulation, and continued existence during adverse environmental conditions
(Bahn et al., 2007; Judelson and Blanco, 2005; Kauserud et al., 2008; Kumamoto,
2008). In response to environmental stimuli, such as heat or drought, fungal
organisms can become “dormant”—an inactive state during which growth and
development cease but from which the organisms can be revived—or transform
into forms that are resilient to heat, drought, and winds. As discussed at the meet-
ing, the environment and environmental stimuli may also serve as a reservoir and
34 FUNGAL DISEASES
trigger for fungal pathogen adaptation and evolution (Casadevall, 2007; Lin and
Heitman, 2006; Stukenbrock and McDonald, 2008).
Sexual reproduction in fungi typically requires the presence of two differ-
ent mating types (Heitman, 2006). Two signals that regulate the sexual cycle of
C. gattii are interactions with plants and extreme desiccation (Lin and Heitman,
2006; Xue et al., 2007). According to Heitman, evidence suggests that when only
one mating type is present in an environment, C. gattii will adopt a “same-sex”
mating strategy for reproduction (Fraser et al., 2005; Lin et al., 2005; Saul et
al., 2008). This adaptability may be a widespread phenomenon, one that enables
recombination, the generation of genetic diversity, and the geographic expansion
of fungi (Heitman, 2006, 2009).52
Same-sex mating may have contributed to the expansion of C. gattii’s geo-
graphical range to Vancouver Island and the U.S. Pacific Northwest (Fraser et al.,
2005). Heitman also discussed how recombination between C. gattii lineages of
the same “sex” may have resulted in a “hypervirulent” recombinant genotype as-
sociated with the outbreak. Two of the three pathogen genotypes associated with
the C. gattii outbreak (VGIIa and VGIIc) are considered highly virulent (Byrnes
et al., 2010; Fraser et al., 2005). Moreover, VGIIa is considerably more virulent
than VGIIa isolates from other parts of the world (Fraser et al., 2005). Although
the reason why the VGIIa and VGIIc genotypes are so virulent is unclear, there
may be a link between the capacity for mating and production of spores and viru-
lence (Byrnes et al., 2010; Fraser et al., 2005; Lin and Heitman, 2006).
Howlett explained that plant breeders consider fungal pathogens to have a
“high evolutionary potential’’ if organisms undergo prolific sexual reproduction
and produce large numbers of genetically diverse spores that then act as inocu-
lum. This capacity leads to frequent breakdowns in a host plant’s resistance to
infection by particular strains of a fungal pathogen. Howlett also explained how
the agricultural environment plays a role in the breakdown of resistance.
Agricultural crops are large swaths of genetically identical plants that “exert
high levels of selection pressure” on populations of fungal strains produced dur-
ing sexual reproduction. Of the billions of offspring produced, the few fungal
strains that can infect these “resistant” plant strains will be amplified with each
subsequent disease cycle. As the frequency of virulent pathogens increases, host
resistance to disease eventually breaks down. Howlett noted that this is exactly
what happened with Leptosphaeria maculans, the causative agent of blackleg in
canola. In 2000, a new cultivar53 of L. maculans with a major resistance gene was
released on the Eyre Peninsula, Australia. Within 3 years, the fungal pathogen had
developed the capacity to overcome the host species’ genetic resistance resulting
in yield losses of more than 90 percent (Sprague et al., 2006).
The environment can also serve as a reservoir for pathogen adaptation and
52╛╛See
contributed manuscript by Heitman in Appendix A (pages 226–248).
53╛╛A
variety of a plant that has been created or selected intentionally and maintained through
cultivation.
WORKSHOP OVERVIEW 35
evolution. Fungal pathogens that are free living in the environment may acquire
what Casadevall called “accidental virulence.” He noted that the soil can be
an extreme environment and that soil-dwelling microbes must adapt to rapidly
changing, often harsh, conditions (Casadevall and Pirofski, 2007). Traits acquired
in this environment, which allow fungal species to survive predation from amoeba
and other protozoan organisms, may also contribute to virulence capabilities in
hosts never before encountered by fungal pathogens. Casadevall suggested that
the concept of “accidental virulence” might best describe how environmentally
acquired fungi can be so virulent in “new” mammalian and other host organisms
(Casadevall, 2007; Casadevall and Pirofski, 2007).
Cryptococcus gattii
FIGURE WO-12╇ Map of the Pacific Northwest, comprising parts of British Columbia,
Figure
Canada, and the states of Washington WO-12.eps
and Oregon in the United States, showing human and
veterinary Cryptococcus gattii cases (including
bitmap mammals) by place of residence or
marine
detection, and locations of environmental isolation of C. gattii during 1999–2008 (strain
NIH444 [Seattle] or CBS7750 [San Francisco] not included). Data were collected from
various state health departments and published reports referenced in the text. The map and
icons have been used at a scale that shows gross geographic areas, effectively masking any
personally identifiable patient locality information. Use of the map is courtesy of exclusive
permission from Google Maps: ©2008 Google, map data ©2008 NAVTEQ.
SOURCE: Datta et al. (2009a).
WORKSHOP OVERVIEW 37
Appendix A, pages 101–116.) Investigators still do not know the origins of the
current epidemic, how C. gattii was introduced into the Pacific Northwest, or how
this invasive fungal pathogen is spreading (Datta et al., 2009a).
Phenomenology
C. gattii is a basidiomycetous yeast that colonizes tree bark, decaying wood,
and nearby soil and is a cause of cryptococcosis, a potentially fatal infection in
humans and animals (Galanis and MacDougall, 2010; Levitz, 1991; Lin and
Heitman, 2006; MacDougall et al., 2007). Before the current outbreak in British
Columbia, Canada and the Pacific Northwest, the environmental source with
which C. gattii had been most often associated was the wood, bark, and detritus
of eucalyptus trees (Levitz, 1991). More recent and widespread global surveil-
lance has established that the fungus also colonizes other tree species (Lin and
Heitman, 2006).54
Individuals become exposed to C. gattii by inhaling the organism or its
spores from soils or trees that have been colonized by the fungus (Lin and
Heitman, 2006; Sorrell, 2001). Once inhaled, C. gattii can cause severe infection
of the lungs and brain, including pneumonia, meningoencephalitis, and cryp-
tococcomas. Unlike C. neoformans, which has become a major cause of death
in HIV-infected individuals around the world, C. gattii also infects apparently
healthy, immunocompetent individuals (Galanis and MacDougall, 2010). The
disease affects a wide variety of humans and animals, but no case of transmission
between animals and/or humans has ever been documented (CDC, 2010; Datta
et al., 2009a).
Timely diagnosis of a C. gattii infection can be difficult. Patients infected
with this fungal pathogen often remain asymptomatic for 6 months or more.
When symptoms do present, fungal agents are not commonly considered by
physicians when evaluating pulmonary disease in an otherwise healthy patient
(Knox, 2010). Treating infected individuals can also be challenging because the
disease tends to require prolonged antifungal therapy, sometimes with multiple
drug courses (Iqbal et al., 2010; Sorrell, 2001). Some have suggested that when
compared with C. neoformans, C. gattii infections tend to require more prolonged
and invasive treatment (Sorrell, 2001). Harris remarked that “existing data sug-
gest that not all cryptococcal infections are alike” and “it is not clear which fac-
tors are the most influential on the patient’s presentation—the species, subtype,
host immune status, or host genetics, or some combination of factors.” 55
54╛╛See
contributed manuscript by Bartlett in Appendix A (pages 101–116).
55╛╛For
more information, see contributed manuscripts by Harris and Heitman in Appendix A (pages
207–225 and 226–248).
38 FUNGAL DISEASES
fence posts, Bartlett remarked. Sampling results have also illustrated the effects
of forestry activities on the abundance of C. gattii in the air (measured in colony-
forming units,56 or CFUs). According to Bartlett, air samples taken in endemic
areas, where trees were being removed, revealed a baseline concentration on the
order of 100 CFU/m3, compared to 10,000 CFU/m3 during tree chainsawing and
wood chipping in the same area (Kidd et al., 2007). Once it is in the air, Bartlett
observed “the organism can travel 10 kilometers, easily, probably further than
that.” The organism is also resilient: Bartlett noted that she can still isolate viable
propagules from sawdust samples taken in 2001.
et al., 2009a). Bartlett remarked that the environmental sampling data may help to
define C. gattii’s ecological niche in British Columbia. The distribution of C. gat-
tii in the environment, thus far, appears heterogeneous, with colonization levels
differing significantly in regions such as the west and east coasts of Vancouver
Island, which, observed Bartlett, are “dramatically different in terms of rainfall,
soil type and vegetation” (Figure WO-13).
Further research into the reasons why C. gattii emerged in the temperate rain-
forest of the Pacific Northwest is needed because it could help researchers predict
the fungus’s future spread and further the scientific community’s understanding of
how environmental pathogens establish themselves in new environmental niches.
Last year we estimate that we found between 10,000 and 20,000 dead bats on
the cave floor€.€.€.€and to be honest the mortality is so disturbing.€.€.€. We just
can’t crawl through so many piles of dead bats.
—Scott Darling, Vermont Fish and Wildlife Department
(Buchen, 2010, p. 144)
FIGURE WO-14╇ Signs of bat white-nose syndrome (WNS). (A) Mortality caused by
Figure WO-14.eps
WNS. (B) Geomyces destructans fungus, forming the visually distinctive white growth
2 bitmaps
on the muzzle, ears, and wings of an infected bat.
SOURCE: Photos provided courtesy of Alan C. Hicks, New York State Department of
Environmental Conservation.
at the U.S. Fish and Wildlife Service fear the extinction of entire New World57
bat populations in the United States and Canada (FWS, 2011).
Such extinctions could have devastating ecological and economic conse-
quences. Bats play a critical role in plant pollination, seed dissemination, and the
control of flying insects, including mosquitoes, moths, beetles, and other night-
flying insect populations (Blehert et al., 2009; Boyles et al., 2011). Large-scale
declines or complete disappearances of bat populations could result in reduced
plant pollination (which is already under siege by colony collapse disorder in
honeybees), significant increases in “nuisance” insect populations, and increased
insect damage to agricultural and forestry resources (Blehert et al., 2009; Boyles
et al., 2011; FWS, 2011). Because bats have very low reproductive rates—pro-
ducing one pup a year, sometimes two, in a single litter—WNS is predicted to
have long-lasting effects (Barclay et al., 2004; FWS, 2011). The value of bats to
the agricultural industry has been estimated to be roughly $22.9 billion/year, with
a range per year of $3.7 billion to $53 billion (Boyles et al., 2011).
57╛╛Refers
to the Western Hemisphere; in a biological context, New World species are those from
the Nearctic Neotropic ecological zones, versus Old World species from the Palearctic and Afrotropic
ecological zones.
WORKSHOP OVERVIEW 43
Phenomenology
Geomyces destructans hyphae58 and conidia59 invade the hair follicles and
sebaceous and sweat glands of bats hibernating in caves and mines with seasonal
temperature ranges between 2°C and 14°C (Blehert et al., 2009). The skin of af-
fected bats does not typically show signs of inflammation or an immune response
at the site of fungal invasion (Meteyer et al., 2009). Hibernating bats infected with
WNS often have severely depleted fat reserves, which are critical for successful
hibernation (Blehert et al., 2009).60
Researchers have not yet confirmed whether Geomyces destructans is the
primary pathogen that causes WNS and the eventual death of affected bats, or if
it is an opportunistic infection that invades animals already immunocompromised
by some other, yet to be defined pathogen (Puechmaille et al., 2010). How WNS
kills bats is also unclear. Infection by the pathogen may irritate the animals, rous-
ing them from hibernation, and causing them to deplete their fat reserves to such
an extent that they are unable to survive through the winter (Blehert et al., 2009;
Cryan et al., 2010; FWS, 2011). Speaker David Blehert of the National Wildlife
Health Center at the U.S. Geological Survey presented recent research suggest-
ing that the fungus causes damage to the wing epidermis and skin structures that
help protect against water loss, causing bats to lose too much water to survive
their winter hibernation (Cryan et al., 2010). (Dr. Blehert’s contribution to the
workshop summary report can be found in Appendix A, pages 167–176.)
a fungus. A large mass of hyphae is known as a mycelium, which is the growing form of most fungi.
59╛╛Asexually produced fungal spore. Most conidia are dispersed by the wind and can endure
extremes of cold, heat, and dryness. When conditions are favorable, they germinate and grow into
hyphae.
60╛╛See contributed manuscript by Blehert in Appendix A (pages 167–176).
44 FUNGAL DISEASES
WO-15 new
WORKSHOP OVERVIEW 45
Figure
FIGURE WO-16╇ Species affected by batWO-16.eps
white-nose syndrome (WNS).
SOURCE: Merlin D. Tuttle, Bat Conservation International, www.batcon.org.
6 bitmaps w vector type
on skin samples from the endangered grey bat (M. grisescens), the southeastern
bat (M. austroriparius) and the cave bat (M. veilfer) (FWS, 2011). Investigators
are concerned that the endangered Virginia big-eared bat (Corynorhinus townsen-
dii virginianus) may also be “at risk.” Although the fungus has been detected on
other bat species that share their hibernacula, there has yet to be confirmed case
of an infected animal (FWS, 2011).
Investigators are concerned that WNS could eventually spread to infect all 25
of the hibernating bat species native to the United States, threatening more than
50 percent of the native U.S. bat populations (Bat Conservation International,
2010). In late May 2010, FWS officials reported that a live bat from Oklahoma
was PCR-positive for G. destructans DNA. This finding alarmed many, because
the infected species, the cave bat M. velifer, frequently shares hibernacula with
other bat species with migratory ranges that extend across the western United
States into Mexico, increasing the potential for further spread of the disease to the
west and south (Bat Conservation International, 2010; Oklahoma Department of
Wildlife Conservation and FWS, 2011). This finding, however, was not confirmed
by fungal culture or histopathology; and, to date, WNS has not been confirmed
in states west of the Mississippi River.
Bat researchers and human cavers may also be contributing to the rapid
spread of this pathogen within and across states. Blehert noted that the index site
for WNS, Howes Cave, is connected to Howe Caverns, a commercial tourist cave
that entertains up to a quarter-million visitors per year. Although there are no data
supporting (or refuting) the hypothesis that humans are serving as transmission
46 FUNGAL DISEASES
“vectors,” Blehert noted that the U.S. Fish and Wildlife Service recommends that
all people who enter caves (e.g., researchers and speleologists) employ decon-
tamination protocols for potentially contaminated clothes and equipment.
Origins
The origins of WNS and its relationship to G. destructans remain unknown
(Qaummen, 2010). However, recent evidence from Europe and characterization
of the newly described pathogen may provide clues. In early 2010, researchers
published several reports confirming that a number of bats (all Myotis spp.) in
France, Hungary, Germany, and Switzerland, while infected with G. destructans,
remained healthy (Puechmaille et al., 2010; Wibbelt et al., 2010). Speaker Gudrun
Wibbelt, of the Leibniz Institute for Zoo and Wildlife Research, reported that not
all of the European hibernacula surveyed have infected bats, and of those that do,
often only a small number of bats within the colony are infected. (Dr. Wibbelt’s
contribution to the workshop summary report can be found in Appendix A, pages
368–403.) Researchers have now confirmed the presence of G. destructans on 8
species of Old World bats in 12 countries in Europe (Puechmaille et al., 2011). 61
In addition, photographic evidence suggests that the fungus was present on bats
in Europe at least as early as the 1980s (Wibbelt et al., 2010). These findings,
which have important implications for future WNS research, could help explain
the origins of the disease and may also provide clues for understanding the
mechanisms of the infection.
One possible interpretation of these data is that G. destructans may have
originated in Europe and that Old World bats and this pathogen may have co-
evolved (Puechmaille et al., 2011; Wibbelt et al., 2010). This hypothesis is sup-
ported anecdotally by reports from routine winter bat surveys in Europe from
the past 30 years. The surveys occasionally noted the presence of a white fungus
similar in appearance to G. destructans on otherwise healthy bats, Wibbelt noted.
If European bats have coevolved with the fungus, they might be able to muster
a sufficient immune response to control and survive infection by G. destructans
(Puechmaille et al., 2010). Some have proposed the possibility that the microbial
flora of bat skin or other abiotic surfaces in European hibernacula “may have
also coevolved to incorporate G. destructans as a non pathogenic component of
the microbial community” (Wibbelt et al., 2010). Wibbelt also reported the pos-
sibility that Old World bats may only be colonized in a superficial fashion on the
outer epidermis without any invasion into deeper tissues.
A second possible interpretation of the discovery of healthy European bats
infected with G. destructans is that disease transmission in Old World bat popu-
lations may be affected for some biological or behavioral reason (Puechmaille
et al., 2010, 2011; Wibbelt et al., 2010). European bats tend to hibernate in
relatively small groups (rarely more than 100 individuals per cluster), Wibbelt
explained. This might make it more difficult for the disease to spread. In the
United States, bats hibernate in groups that can reach into the hundreds of thou-
sands (Wibbelt et al., 2010).
Additional explanations for why European bats infected with G. destructans
do not succumb to WNS include the possibility that the fungal strain in the United
States is more virulent than the strain in Europe or that G. destructans is not the
primary cause of death in WNS. However, Blehert did note that his team’s and
others’ diagnostic investigations of infected New World bats had ruled out tox-
ins, parasites, and known viral and bacterial pathogens as associated with WNS
(Blehert et al., 2009; Chaturvedi et al., 2010; Gargas et al., 2009). Wibbelt ob-
served that G. destructans isolates from North America and Europe appear identi-
cal in morphology and in the sequence of two genes (ITS and SSU) commonly
used as a marker to distinguish between different species (Puechmaille et al.,
2011). Further research will be necessary to determine the true cause of differ-
ences between North American and European bats infected with G. destructans.
Amphibian Chytridiomycosis
Phenomenology
B. dendrobatidis is an aquatic chytrid fungus—an early diverging class of
fungi—that infects keratinized epidermal cells of amphibians, causing rapidly
progressing and deadly chytridiomycosis in susceptible species (Fisher et al.,
2009). First identified in 1998 and characterized in 1999, Bd is unique among
other chytrids (Berger et al., 1998; James et al., 2006; Longcore et al., 1999).
It is one of only two known chytrid fungi to parasitize vertebrates and the only
known species to infect the keratinized skin of living amphibians (Berger et al.,
1998; Fisher et al., 2009). Bd’s asexual spore is a “free living” and motile zoo-
spore, possessing a single flagellum that allows the spore to travel small distances
(usually less than 2 cm) and thrive in aquatic habitats such as streams and ponds
(Fisher et al., 2009; Kilpatrick et al., 2009; Kriger and Hero, 2007; Rosenblum
et al., 2010).
According to speaker Vance Vredenburg, from San Francisco State Univer-
sity, investigators believe that amphibians become infected with Bd through both
casual contacts with zoospores in the water as well as through direct animal-
to-animal transmission. Once infected, the susceptible animals carry Bd for
24 to 220 days before they succumb to chytridiomycosis (Lips et al., 2006).
Vredenburg remarked that the susceptibility to Bd colonization and subsequent
development of chytridiomycosis varies widely across species, populations, and
individuals. Indeed, laboratory experiments have found mortality rates from 0 to
100 percent, depending on temperature, species, and age of the infected animals
(Berger et al., 2005; Daszak et al., 2004; Kilpatrick et al., 2009; Lamirande and
Nichols, 2002; Woodhams et al., 2003).
Many of the most notable and rapid declines in amphibia have occurred
among those populations living at high altitudes in mountainous regions, leading
some to associate outbreaks of fatal chytridiomycosis with cooler climates and
high altitudes (see Fisher et al., 2009). Several amphibian species that live near
sea level also have experienced notable population declines from this fungal dis-
ease, however, suggesting that this association may be an “oversimplification of
a complex host–pathogen relationship” (Fisher et al., 2009). The preponderance
of the evidence supports the observation that Bd does prefer cooler temperatures,
growing and reproducing between 4°C and 25°C (Berger et al., 2004; Drew
et al., 2006; Kilpatrick et al., 2009; Kriger and Hero, 2007). Indeed, the available
evidence suggests that the virulence of Bd is inversely related to temperature,
perhaps in a species-dependent manner (Fisher et al., 2009; Walker et al., 2010).
The specific mechanisms by which some species suffer rapid declines when
Bd is introduced while others are able to tolerate varying levels of infection
without the development of disease—or even resist infection entirely—remain
50 FUNGAL DISEASES
ated with lethal infections. Bd may disrupt the normal regulatory functions of am-
phibian skin, causing osmotic imbalances, electrolyte depletion, and ultimately
death (Rosenblum et al., 2010). Although some investigators have suggested that
Bd might release lethal toxins, no specific toxin has been identified (Fisher et al.,
2009; Rosenblum et al., 2010). As discussed by Vredenburg, investigators are
also exploring the possibility that other microbial components of the amphibian
skin microbiome can contribute to disease mitigation (Harris et al., 2009).
The first reports of sudden oak death occurred in California forests in 1994–
1995 (Rizzo and Garbelotto, 2003). While the origins of the associated pathogen,
Phytophthora ramorum (P. ramorum), remain unknown, investigators believe the
“source” of the subsequent epidemic in California was an infected ornamental
Rhododendron plant(s) (Kliejunas, 2010; Mascheretti et al., 2009). Now known
to cause disease (commonly known as ramorum blight) in more than 100 plant
species, P. ramorum has also emerged as a novel plant pathogen in the United
Kingdom and Europe (Grünwald et al., 2008). Collectively, these diseases have
led to the rapid decline of oak forests on the west coast of the United States and
have led to widespread disease in trees and woody ornamental plants throughout
the United Kingdom and Europe (Grünwald et al., 2008).
P. ramorum is just one of the many invasive tree diseases that were intro-
duced into the forests of North America in the 20th century. As in the case of
chestnut blight and Dutch elm disease, the transportation and trade of plants and
plant materials contributed to the movement of this pathogen across oceans and
continents, as well as between suburban and forest ecosystems (Brasier and Web-
ber, 2010; Goss et al., 2009; IOM, 2010). As reviewed by speaker Rizzo, once P.
ramorum is established in the landscape, treatment options are extremely limited.
The discovery in 2009 that P. ramorum was reproducing in Japanese larch trees
in the United Kingdom led to the immediate clear cutting of 4 million larch
trees—more than 10,000 acres of forest—in a heroic effort to slow the spread of
the disease (Hardman, 2011).
Phenomenology
P. ramorum is a fungus-like oomycete63 or “water mold” that thrives in
the cool, wet climate of California coastal forests (Kliejunas, 2010). In con-
trast to most species of Phytophthora, P. ramorum exhibits a remarkably broad
63╛╛As
noted by speaker Rizzo, Phytophthora spp. is not a “true fungus.” It is an oomycete or “water
mold” that belongs to the Kingdom Stramenopila (a major eukaryotic group that includes diatoms
and brown algae, and is distinct from plants, fungi, and animals). Like fungi, oomycetes “exhibit fila-
mentous growth, produce sexual and asexual spores, and can feed on decaying matter or be obligate
parasites of plants” (Kliejunas, 2010).
54 FUNGAL DISEASES
host range, infecting a variety of tree and non-tree species—ranging from hard-
wood and conifer trees, to shrubs and leafy plants64 (Brasier and Webber, 2010;
Grünwald et al., 2008). P. ramorum colonizes the leaves of many plants and the
inner bark and sapwood of trees. The fungus can also survive in a dormant state
in decaying matter, such as leaf litter on the forest floor, and in the soil (Parke and
Lucas, 2008). Rizzo remarked that within forest ecosystems in California, “once
we started looking, we started finding it on just about every plant we looked at,
ranging from ferns to redwood trees.”
Infection occurs when spores and zoospores—dispersed by winds or wa-
ter—come into contact with susceptible plants. Moisture is not only essential
for the production of infectious propagules, but free water on plants—from fog,
dew, or rainfall—enhances infection and dispersal (Kliejunas, 2010; Rizzo and
Garbelotto, 2003). Indeed, monitoring streams that are “baited” with Rhododen-
dron leaves, Rizzo said, has been an effective method for early detection of this
pathogen in forest ecosystems. Disease manifests differently depending on the
plant species and the part of the host plant that is infected (Grünwald et al., 2008)
(Figure WO-19 and Table WO-2). Bleeding lesions and stem cankers develop
in forest trees, followed by rapid declines and “sudden” death. Ramorum blight
causes shoot-tip dieback and leaf spots in woody shrubs and ornamental trees.
Unlike sudden oak death, ramorum blight rarely kills its host (Grünwald et al.,
2008).
A B
FIGURE WO-19╇ Sudden oak death and ramorum blight. (A) Aerial view of a forest in
Humbolt County with patches of trees dying of sudden oak death; (B) canker on tanoak;
(C) signs of ramorum blight on a variety of host plant species.
SOURCE: Rizzo (2010). WO-19 New
56 FUNGAL DISEASES
Southern California plant nursery, but not before the nursery had shipped poten-
tially infected plants to more than 40 states (Figure WO-20) (Goss et al., 2009).
As Rizzo noted, many of the fungicides used in the nursery trade do not kill
the pathogen, but rather suppress symptoms; this is probably how P. ramorum
spreads over very long distances via the plant trade. Leaves, flowers, and stems
of infected plants carry the pathogen, which is also spread by the transporta-
tion of plant or associated plant materials, including soil (Kliejunas, 2010).
Waterways are an effective means of spreading P. ramorum. Investigators have
detected the pathogen in streams contaminated with run-off from infected nurser-
ies (Kliejunas, 2010). P. ramorum may also provide an interesting case study of
humans acting as vectors for fungal disease, as sudden oak death “has potentially
been spread to new areas by hikers, mountain bikers, and equestrians” (IOM,
2010). Rizzo also pointed to the movement of “green waste,” that is, compost,
firewood, mulch, and other plant matter, as an another possible means of spread-
ing the fungus within and across ecosystems.
As of 2009, the U.S. Department of Agriculture (USDA) Animal and Plant
Health Inspection Service reported detection of P. ramorum in 11 states (Ala-
bama, California, Georgia, Maryland, Mississippi, New Jersey, North Carolina,
Oregon, Pennsylvania, South Carolina, and Washington) at 30 sites (24 nurseries
and 6 in the landscape) (Kliejunas, 2010). The pathogen’s eastward spread has
placed Eastern native forests at risk and has led many experts to worry that P.
ramorum could have ecological consequences comparable to those of chestnut
blight and Dutch elm disease (Goss et al., 2009).
The distribution of sudden oak death in California forests is heterogeneous,
Rizzo noted. Weather—including winds, temperature, humidity—contribute to
the pathogen’s establishment and spread within a new environment (Kliejunas,
WORKSHOP OVERVIEW 57
FIGURE WO-20╇ P. ramorum “migration” pathways. The ornamental plant trade contin-
ues to serve as a major pathwayFigure WO-20.eps
for the spread of P. ramorum. Arrows indicate confirmed
P. ramorum-positive nursery trace forwards. Blue arrows are 2004 trace forwards and red
bitmap
arrows are 2006 trace forwards. There were no confirmed trace forwards in 2005 or 2007.
Pie charts show the distribution of the two groups of isolates (of NA 1 lineage) among
sampled states.
SOURCE: Goss et al. (2009).
Origins
The lack of reports of P. ramorum in the United States before the mid-
1990s, combined with its aggressiveness and limited geographic range relative
to its hosts’ distribution, suggests that P. ramorum was only recently introduced
58 FUNGAL DISEASES
into the United States (Grünwald et al., 2008). Genetic evidence supports the
hypothesis that the U.S. and European strains are distinct and that both strains
likely originated from a third, as yet unknown, source (Grünwald et al., 2008).
Molecular studies have identified three main lineages: (1) EU1—found in both
North American and European nurseries and some European woodlands; (2)
NA2—found in California and Washington nurseries; and (3) NA1—found in
North American nurseries and in California and Oregon forests (Grünwald et al.,
2008).
An alternative explanation is that P. ramorum may have existed in California
for many years, but only recently emerged because of changes in the environment
(e.g., increased temperatures, fire suppression, modifications in land use patterns)
that have led to an increased prevalence and aggressiveness of the pathogen
(Rizzo and Garbelotto, 2003). Native P. ramorum or other Phytophthora sp. may
also have evolved into a more virulent form, may have undergone a change in
host specificity or preference, or may represent an entirely new hybrid species
(Rizzo and Garbelotto, 2003). It has been reported that a novel Phytophthora
hybrid—a cross between P. cambivora, an oak pathogen, and P. fragariae-like
isolates, a strawberry pathogen—emerged in Europe in the 1990s and has killed
thousands of alder trees (Alnus spp.) (Brasier et al., 1999). Given the large num-
ber of Phytophthora spp. in California agricultural and horticultural environ-
ments, a hybrid origin for P. ramorum is certainly a possibility; although other
explanations are also possible (Rizzo and Garbelotto, 2003; Tyler et al., 2006).
The presence of two virulent and highly aggressive yellow rust strains at high
frequencies at epidemic sites on five continents may represent the most rapid and
expansive spread ever of an important crop pathogen. This epidemic trend may
continue because the aggressive strains, which can tolerate higher temperatures,
are still evolving.
—Hovmøller et al. (2010, p. 369)
65╛╛Speaker Mogens Hovmøller defined “aggressiveness” as the quantitative ability to cause more
FIGURE WO-21╇ Wheat production regions worldwide. Each red dot represents 20,000
tons of wheat production.
Figure WO-21.eps
SOURCE: Trethowan et al. (2005). 3 bitmaps
and Eastern Africa, Western and Central Asia, and China (Hovmøller et al., 2010).
According to Hovmøller et al. (2010), this may be the most rapid and expansive
spread of an important crop pathogen ever documented.
Wheat is the most widely grown cereal crop, produced as food for humans
and as feed for livestock (Figure WO-21). Worldwide, wheat accounts for one
fifth of the total human caloric intake. In regions such as Western Asia, it can
account for as much as half of the daily calorie intake (Hovmøller, this volume;
Stone, 2010). Epidemics of rust disease have been held in check by rust-resistant
wheat cultivars developed in the mid-20th century, and other agricultural prac-
tices (Koerner, 2010). Since 2000, however, the natural history66 of yellow rust
appears to have changed. Disease is now emerging in regions previously consid-
ered inhospitable to this fungal pathogen (Hovmøller et al., 2008; Milus et al.,
2009).
Phenomenology
Puccinia striiformis f. sp. tritici (hereafter, P. striiformis) is an obligate,
basidomycetous, pathogen of wheat. It infects the green tissues of host plants,
causing damage to leaf blades, and reducing the yield and quality of produced
grains and seeds (Chen, 2005). Severe infections of yellow rust also stunt the
growth of wheat plants. Average reported yield losses due to yellow rust range
from 10 to 70 percent; yield losses in highly susceptible cultivars can reach 100
percent (Chen, 2005). Weather—including humidity, rainfall, temperature, and
66╛╛The natural development of something (as an organism or disease) over a period of time.
60 FUNGAL DISEASES
wind—is critical for developing favorable conditions for fungal infection and
growth (Chen, 2005).
Named for the yellow pustules of powdery spores (urediniospores) that ap-
pear as “stripes” on the leaf blades of infected plants, yellow “stripe” rust is an
important “cooler climate” wheat disease (Chen, 2005) (Figure WO-22). Disease
can occur early in the growing season, when temperatures are low. Areas affected
by this disease tend to be in temperate regions and high-elevation areas in the
tropics (Chen, 2005). P. striiformis depends on a living host for survival and
produces huge numbers of airborne spores that are carried by wind from one
susceptible host to another (Brown and Hovmøller, 2002).
bitmap
62 FUNGAL DISEASES
MT ND
MN
ME
Before 2000 OR ID SD WI
MI
V
WY NH
NY TMA
IA
NE PA CT
NV UT IL IN OH MDNJ
CO KS MO WV DE
KY VA
CA OK TN NC
AZ NM AR SC
LAMSAL GA
TX
Trace – 20% FL
Severe (>20%)
WA MT ND
2001 and thereafter MN ME
SD WI VT
OR ID MI
WY IA NY NH
NE IA MA
NE PA CT
NV UT CO ILIN OH MD NJ
IL
KS
KS MO
MO WV DE
KY VA
CA OK AR TN NC
AZ NM AR NC
SC
MS AL GA
MSAL
TX LA
FL
FIGURE WO-23╇ Presence of “trace” and “severe” levels of yellow rust in North America
since 2000.
SOURCE: Adapted from Chen (2005).
Figure WO-23.eps
of these strains and may explain their rapid spread on a global scale (Hovmøller
et al., 2008). In Australia and the United States, these newer, more aggressive
strains appear to have replaced older strains and have continued to thrive in sub-
sequent seasons (Milus et al., 2009). According to Hovmøller, while the epidemic
has apparently slowed down in North America from 2006 to 2009, major yellow
rust epidemics occurred in northern and eastern Africa, Western and Central
Asia, China, and the Middle East (Hovmøller et al., 2010). In 2010, according to
Hovmøller, observers reported disease outbreaks in northern and eastern Africa,
Asia, and the Middle East, and the epidemics reappeared in the United States.
Origins
The exact geographic origin of these more aggressive strains of P. striiformis
is unknown, although phylogenetic analyses reveal that West and Central Asia
may be the evolutionary origin. Hovmøller remarked that the dramatic change in
WORKSHOP OVERVIEW 63
phenotype (e.g., spore production, heat tolerance), coupled with the sudden ap-
pearance of these strains suggests a recombination event somewhere, rather than
evolution through a series of mutations.
Unlike many of the fungal diseases discussed at the workshop, yellow rust
is well known. Epidemics of this disease have plagued the world’s farmers for
centuries, and P. striiformis is familiar to plant breeders. In the 1940s, Norman
Borlaug68 developed new “rust-resistant” wheat strains that also dramatically
increased global crop yields (Rust in the bread basket, 2010). These advances in-
spired the “Green Revolution” that brought these techniques and disease-resistant
wheat into widespread use. According to Brown and Hovmøller (2002), these
techniques have helped to keep many crop diseases under control, but now,
relatively few crop varieties (with specific resistance genes) are sometimes used
across large areas at “continental scales.” A reduced crop diversity in modern
agriculture, according to Hovmøller, increases the potential impact posed by the
global movement of new, more virulent forms of plant pathogens (Brown and
Hovmøller, 2002; Stukenbrock and McDonald, 2008).69
68╛╛Norman Borlaug was an American agronomist. His work to develop disease-resistant crop strains
earned him the titles of Nobel laureate and “the father of the Green Revolution.” Borlaug was one
of only six people to have won the Nobel Peace Prize, the Presidential Medal of Freedom, and the
Congressional Gold Medal. He was also a recipient of the Padma Vibhushan, India’s second highest
civilian honor.
69╛╛In 1999 a new, highly virulent, strain of wheat stem “black” rust, Ug99, emerged in Uganda.
This pathogen was able to circumvent the genetic resistance of wheat hybrids developed during the
Green Revolution. This pathogen has now spread to Kenya, Ethiopia, Sudan, Yemen, and Iran (Vurro
et al., 2010). Researchers worry that Ug99 will soon spread to the major wheat growing regions of
Pakistan and India, which account for ~20 percent of the world’s wheat supply (Vurro et al., 2010).
64 FUNGAL DISEASES
is too huge, too unknown, and too threatening not to develop improved capacity
for detection, diagnosis, and response to emerging fungal pathogens and diseases.
David Blehert, joined by Forum member Russell and fellow Forum member
Jacqueline Fletcher of Oklahoma State University, emphasized the importance of
cross-disciplinary communication and a “One Health”70 approach for developing
a more robust capacity for global disease surveillance, detection, and response.
As noted by Bartlett, the detection and response to the emergence of C. gattii in
British Columbia and the U.S. Pacific Northwest in 1999 ultimately involved pro-
fessionals and expertise from the veterinary, medical, public health, and plant and
wildlife communities. Indeed, because of its close interactions with plants and
ability to cause disease in humans and animals, C. gattii was called the “poster
child” for a One Health approach.
The plant health community was considered by many at the meeting to be
an equal partner in a One Health approach to infectious disease detection and
response. Fungi can form associations with and, in some cases, be pathogenic to
humans, animals, and plants, Fletcher observed. She described the need to gain
a better understanding of how fungal pathogens might “jump” between plant and
animal or human systems. Others remarked on the indirect, but potentially sig-
nificant, impacts of plant pathogens on human health and well-being, including
threats to ecosystem stability or global food security. As Fletcher noted, however,
“plant pathogens have only been incorporated into current One Health initiatives
[in] a very minor way.”
Surveillance Networks
Over the past several decades, various systems for passive and active sur-
veillance for emerging and reemerging diseases of humans, animals, and plants
have been developed at regional, national, and international levels (GAO, 2010).
Systems supporting the surveillance and detection of emerging fungal plant
pathogens are the most sophisticated, possibly due to the historical importance
of fungal diseases on economically important foodstuffs, crops, and plants (IOM,
2007; Rossman, 2009).
Disease surveillance and detection in the United States is a shared responsi-
bility of various state and federal programs (Figure WO-24) (GAO, 2010). The
Centers for Disease Control and Prevention (CDC) and other federal agencies, in-
cluding the USDA, Department of Defense (DoD), and Department of the Interior
(DOI), independently gather and analyze national infectious disease surveillance
reports as well as morbidity and mortality data for humans, plants, livestock, and
70╛╛One
World, One Health® is a registered trademark of the Wildlife Conservation Society and
reflects the need to establish a more holistic approach to preventing epizootic disease and for main-
taining ecosystem integrity for the benefit of humans, their domesticated animals, and the foundation
biodiversity that supports us all. For more information, see http://www.oneworldonehealth.org/ (ac-
cessed April 11, 2011).
WORKSHOP OVERVIEW 65
FIGURE WO-24╇ Roles and responsibilities for monitoring for pathogens in humans,
animals, plants, food, and the environment in the United States.
SOURCE: GAO (2010). Figure WO-24.eps
bitmap
wildlife. The CDC, USDA, DoD, and DOI independently fund and maintain both
domestic and international laboratory networks for infectious disease diagnostics
(Choffnes, 2008; GAO, 2010). Reporting and verification of outbreaks of specific
diseases of concern or of unusual symptoms or health disturbances takes place at
the state level. The findings are sent to federal agencies for further investigation,
and if appropriate, the coordination of a response to the potential disease threat
(GAO, 2010).
The DOI includes the U.S. Geological Survey (USGS) National Wildlife
Health Center (NWHC), which is tasked with providing information, technical
assistance, research, education, and leadership on national and international wild-
life health issues. According to Blehert, WNS surveillance is currently largely
opportunistic, based upon people making unusual observations in the field, then
66 FUNGAL DISEASES
sending bats to the NWHC or other laboratory for diagnostic investigation. Fed-
eral, state, and local government officials have taken several steps to try to curb
the spread of WNS and prevent additional bat deaths. In 2009, the U.S Fish and
Wildlife Service issued a cave advisory that established guidelines for entering
bat hibernacula, that issued recommendations for decontamination of caving gear,
and that asked researchers and spelunkers not to bring clothing or equipment that
has been used in caves from affected areas to caves in unaffected areas (FWS,
2011). Federal and state agencies have also closed caves on public lands in order
to prevent people from inadvertently spreading the fungus to new areas (FWS,
2011). Total funding for WNS research from federal and state agencies increased
from approximately $1.8 million to $10 million between fiscal year 2007 and
fiscal year 2010.
Formal global surveillance programs—tracking the emergence and reemer-
gence of microbial threats to human, animal, and plant health—are coordinated
by the World Health Organization (WHO), the World Organisation for Animal
Health (OIE), and the Food and Agriculture Organization (IOM, 2007). WHO
manages the Global Outbreak Alert and Response Network which partners with
120 “informal” (as discussed below) and “formal” (e.g., regional WHO offices,
government, military, or university research centers) information sources to iden-
tify and respond to disease outbreaks (Heymann and Rodier, 2004). OIE manages
an international reporting system on animal disease that includes reporting of
“exceptional epidemiological events” and periodic gathering of animal health in-
formation (Jebara, 2004). Forum member Peter Daszak of the EcoHealth Alliance
noted that the responsibility for infectious disease surveillance of wildlife could
be undertaken by the United Nation’s International Union for the Conservation
of Nature (IUCN). The IUCN has developed a working group of 400 wildlife
specialists from around the world.71
competent veterinary authority to the OIE as required under Chapter 1.1 of the Aquatic Code. OIE
members are also required to report the presence or absence of each disease in their territory on a
semi-annual basis, and ensure disease surveillance programs are implemented to support any claims
of freedom from one or both diseases” (Schloegel et al., 2010, p. 4).
WORKSHOP OVERVIEW 67
ing requirement only applies to OIE member states, and only to animals traded
internationally. Indeed, the effectiveness of these formal and informal reporting
regimes has yet to be demonstrated, and many have suggested that fear of ad-
verse economic consequences (e.g., trade and tourism restrictions) will limit their
usefulness as an early warning disease reporting network (Cash and Narasimhan,
2000; Fidler, in IOM, 2010; Hueston, in IOM, 2007; Perrings et al., 2010).
Guidelines on hygiene and quarantine procedures for captive and wild ani-
mals have also been developed by the conservation community to reduce the
spread of zoonotic diseases, but these guidelines are considered underused or dif-
ficult to enforce (Daszak et al., 2000). Two conventions developed to address the
international wildlife trade and the conservation of biodiversity, the Convention
for International Trade of Endangered Species and the Convention on Biologi-
cal Diversity are based entirely on voluntary agreement, noted speaker Weldon.
While these agreements reflect noble aspirational goals, according to Weldon,
there is limited opportunity to actually implement the measures.
Several participants emphasized the limitations of current capacity to detect
emerging pathogenic fungi. As speaker Fisher observed, national strategies are
limited by their focus on known threats to humans and agriculturally important
species, and international strategies are nearly nonexistent or very slow moving.
This is particularly true of wildlife surveillance, which Fisher said is “completely
under the radar.” Forum member Roger Breeze of Lawrence Livermore National
Laboratory agreed, noting that “we are not very good at looking for things we
know about, even those diseases that are economically important, such as foot
and mouth disease.” Breeze continued that “what we are talking about over the
last few days is broadening the number of organisms involved and the number of
areas of economic life that are involved.” He went on to note that many organisms
discussed during the workshop, such as ornamental plants, do not currently fall
under any one organization’s regulatory responsibility. “We have a huge interna-
tional failure in biosecurity,” according to Breeze, and the problem “needs to be
approached in a different manner.”
No single agency or multilateral organization is solely focused on infectious
diseases in humans, plants, and animals. Several workshop participants observed
that the creation of a single entity that was responsible for collecting and analyz-
ing data from across the “threat” spectrum and ensuring that disease interventions
are based on the input of professionals working with humans, domestic animals,
and wildlife could significantly enhance current disease surveillance and response
capabilities (Choffnes, 2008; GAO, 2010; Hueston, in IOM, 2007; Perrings et al.,
2010).
Improved coordination of disease surveillance and response activities,
Daszak noted, would “benefit all sectors—whether it is food production, travel
and trade, or human and environmental health.” Moreover, sectors may benefit in
unanticipated ways. Blehert remarked that “wildlife health is important to world
health. Not just with regard to disease surveillance, but also with regard to basic
68 FUNGAL DISEASES
research. There is much that we can learn from emerging diseases of wildlife such
as WNS or amphibian chytridiomycosis that likely have significant implications
with regard to ecosystem integrity and function. Only by incorporating domestic
animal health, wildlife health, and human health into the same model can we fully
understand the ecology of infectious disease.”
Several participants suggested that within the United States, an interagency
task force could link together the plant, animal, and human health communities.
Forum member Russell added that the Department of Defense is now one of
several interagency partners involved in a forum on emerging pandemic threats
as a sub-Interagency Policy Committee (IPC) of the U.S. government’s Global
Health Initiative (GHI). Among its other activities, this sub-IPC assembled an
interagency working group that developed a document detailing the U.S. response
to the revised IHRs. He suggested that this interagency forum could serve as an
effective model for coordinating the U.S. government activities in areas of com-
mon concern.
Forum member Edward McSweegan, from the National Institutes of Allergy
and Infectious Diseases, added that a previous interagency program that was
focused on international infectious diseases was orchestrated by the U.S. Depart-
ment of State and the Office of Science and Technology Policy (OSTP). Forum
Vice-Chair, James Hughes of Emory University, noted that this effort was estab-
lished under the aegis of the Committee on International Science, Engineering,
and Technology Policy of President Clinton’s National Science and Technology
Council and involved many agencies: the National Institutes of Health (NIH),
CDC, Food and Drug Administration, U.S. Agency for International Develop-
ment, and DoD, among others. He said that many of the recommendations from
their 1995 report are “still relevant to today’s world.”73 McSweegan also sug-
gested that the funding of cross-disciplinary research on emerging fungal diseases
might be modeled after the success of the NIH–National Science Foundation
(NSF) Ecology of Infectious Diseases Initiative, 74 perhaps as a collaboration of
the NIH, USGS, and USDA.
website http://www.fic.nih.gov/programs/research_grants/ecology/index.htm.
WORKSHOP OVERVIEW 69
75╛╛See http://www.promedmail.org.
76╛╛Including clinician reports, blogs, chat rooms, websites, news media, YouTube videos, and other
Internet sources.
77╛╛See at www.bd-maps.net.
70 FUNGAL DISEASES
reported about 6,500 Bd-positive animals out of 30,000 sampled among 3,500
sites worldwide, with 49 of 74 countries and 440 species of amphibians with
known Bd infections. The data come from multiple sources, including contribu-
tions directly from the field using a smart phone application called EpiCollect. 78
Fisher observed that these data may be used to assess either global or country-
level trends and to detect broad-level associations (e.g., the data show that Bd is
present in many areas where species richness has declined without any [other]
explanation). Fisher noted that the project has provided a means for communicat-
ing important information rapidly among interested parties.
Predictive Modeling
Surveillance and response efforts could be better targeted to at-risk popula-
tions or circumstances through the use of mathematical models and Geographical
Information Systems (Weinberg, 2005). Several workshop participants described
the use of predictive modeling as a way to “get ahead” of the spread of invasive
fungal diseases into new and highly susceptible regions:
Daszak discussed the use of predictive models to get ahead of disease emer-
gence entirely by anticipating where viral pathogens of zoonotic origin are most
likely to emerge in the future. The PREDICT project is part of the U.S. Agency
for International Development Emerging Pandemic Threat program. 79 PREDICT
uses wildlife surveillance data and models to identify (1) geographic hot spots for
the emergence of infectious disease, and (2) species that may serve as reservoirs
of disease.80 Daszak noted that the prediction models developed for the PREDICT
project were grounded in wildlife surveillance data that included the active col-
lection of tens of thousands of samples from wildlife among 24 countries. Any
newly discovered viruses in these samples are deemed “high priority” if they ap-
pear to be closely related to other known viral pathogens. High-priority pathogens
are further characterized and, if appropriate, people who interact with the wild-
life that may be affected by these pathogens are educated and advised to avoid
contact. While PREDICT is currently focused only on viral pathogens, Daszak
observed that the same approach could be used for fungi and fungal pathogens.
Harris observed that labs with mycology capacity are not as common as labs with
viral or bacterial capacity, and requests for additional training or capacity need
to come from the local level. She noted that the C. gattii Public Health Working
Group, formed in 2008 by the CDC and state and local public health depart-
ments and laboratories and the British Columbian Centre for Disease Control,
is working to standardize surveillance by increasing clinician awareness of C.
81╛╛During a canavanine-glycine-bromothymol blue agar test, C. gattii grows and the medium
changes color; C. neoformans does not grow, and the medium remains a light green color.
82╛╛Bartlett pointed out that because of the close evolutionary relationship between the two species,
much of the early literature on the outbreak refers to the outbreak pathogen by its former name, C.
neoformans var. gattii.
74 FUNGAL DISEASES
gattii infection and working with global laboratories to characterize genetic and
phenotypic variety in C. gattii.
Padilla, from the Smithsonian Conservation Biology Institute (SCBI), iden-
tified fungal diagnostic capacity as a particularly challenging area of emerging
fungal disease threat management in wild animals and one in need of further
research. Too often, new fungal pathogens in wildlife are either misdiagnosed or
undiagnosed. The limited diagnostic capacity leads to the “clumping” of infor-
mation under known fungal disease syndromes, Padilla remarked, and this often
precludes the recognition of true emerging fungal diseases and prevents further
investigation of true host–pathogen dynamics. Fungal infections are often iden-
tified only to the genus, not the species level, making it difficult to understand
host–species relationships. The clumping of information could also result in
dangerous management decisions—when assumptions about one host are based
on what is known about another host. For example, he reported that preliminary
findings by researchers working at the SCBI suggested that the bacterium Janthi-
nobacterium lividum does not provide the same anti-chytridiomycosis protection
in Bd-infected captive-bred Panamanian golden frogs (Atelopus zeteki), as it does
in Bd-infected yellow mountain frogs (Rana muscosa). In fact, J. lividum seems
limited in its ability to colonize the skin of A. zeteki and thus is also limited in
its ability to play the same anti-fungal role that it does in R. muscosa. However,
the understanding of a bacterium–host system conferring anti-fungal protective
properties suggests that other species-specific host-adapted bacteria could confer
the same protection to A. zeteki. Padilla expressed hope that these findings will
lead to more appropriate treatments for this particular frog species.
84╛╛The study of the general principles of scientific classification, and the classification of organisms
Developing strains of wheat that are resistant to the newly emerging and
more aggressive forms of yellow rust is the primary strategy for limiting the
devastating effects of P. striiformis on wheat. In the meantime, early detection is
essential to reduce crop yield losses due to yellow rust, Hovmøller said. In the
short term, options for control are limited to fungicide sprays which may be un-
available or not affordable to farmers in the developing world. The replacement
of susceptible wheat with locally adapted, resistant, or less susceptible varieties
can also slow disease spread, he remarked.
As Hovmøller noted, when it comes to wheat rust, “what’s going on in
one continent may be your problem the following day.” To prevent long-term
damage, intensified international collaboration is needed to build wheat rust
surveillance, detection, and response capacity. Several promising developments
on the international scale were reported by Hovmøller. In 2008, a Global Rust
Reference Center (GRRC) for yellow rust was established to improve yellow rust
management in countries where facilities and expertise are scarce. GRRC is sup-
ported by Aarhus University in Denmark, the International Center for Agricultural
Research in the Dry Area, and the International Maize and Wheat Improvement
Center (CIMMYT). In 2011 the activities will be extended to wheat stem rust
(Puccinia graminis) via projects facilitated by the Borlaug Global Rust Initiative.
GRRC is complementing existing national diagnostic laboratories, which cannot
receive rust samples year round from all countries. The primary goals of GRRC
are to conduct virulence and race85 analyses, secure isolates for future resistance
breeding and research, facilitate research and training, and provide information
for a global wheat rust early warning system. The Borlaug Global Rust Initiative,
which was established in response to the stem rust Ug99 outbreak in East Africa,
now deals with all three wheat rusts.86
Disease management is considerably more difficult when dealing with plants
of limited economic value (i.e., non-timber and non-crop plants), despite their
significant ecological value, Rizzo noted. Management of sudden oak death
and ramorum blight, according to Rizzo, is “scale dependent.” One can manage
individual trees, landscapes, or entire regions (Rizzo et al., 2005). At the indi-
vidual level, fungicides are available that can be injected into a tree or sprayed
on the bark to prevent infection (Garbelotto et al., 2002). In forests, containment,
including cutting and controlled burnings in areas with infected trees, is the pri-
mary means of infection control. When asked whether there is a possibility for
developing treatments for sudden oak death that could be applied at the landscape
level, Rizzo said the options are limited. In Oregon, there have been attempts to
conduct aerial spraying with phosphonate (a chemical fungicide), but it is un-
likely that any type of aerial spraying would be acceptable in California. Rizzo
also cautioned that it took many decades to breed genetic resistance for Dutch
85╛╛A subspecies group of pathogens that infect a given set of plant varieties (Cornell University,
plant pathology glossary).
86╛╛For more information see www.globalrust.org.
WORKSHOP OVERVIEW 77
elm disease. The complexity of oak genetics also makes it very challenging to
use breeding to develop oaks resistant to P. ramorum infection. Rizzo went on
to note that research on potential biocontrol agents, such as viruses, is at a very
early stage.
Landscape-level management also uses predictive modeling. For sudden oak
death, models based on host species distribution, climate, and other factors iden-
tify areas at risk for invasion by the pathogen. These areas can then be surveyed
using “aerial imaging, plot-based monitoring, and stream sampling to determine
the presence of P. ramorum or signs of infected trees” (IOM, 2008b). Eradication
methods are only effective if the disease is detected early enough. For areas where
the pathogen is established, management approaches seek to avoid negative eco-
logical consequences, such as the growth of invasive plant species. Ultimately,
Rizzo said, we are trying to develop methods to “live with the pathogen.”
87╛╛Valley fever is often dismissed as a self-resolving mild illness. In fact valley fever can be long-
lasting and have a tremendous impact on activity levels (see Galgiani, 2007). Recent surveillance
activities conducted by the Arizona Department of Health Services, in collaboration with the CDC,
reported that coccidioidal illness lasted an average of 6 months, with 75 percent of workers taking
more than one month of sick leave and 40 percent of infected persons requiring at least one night of
hospitalization at some point during the course of their illness (Tsang et al., 2010). Annual hospital
costs alone amount to nearly $90 million ($86 million in 2007; Tsang et al., 2010).
78 FUNGAL DISEASES
University of Arizona acquired the compound in 2005, have clinical trials re-
sumed (Galgiani, 2007).88
In the absence of a vaccine or other preventive measures for C. gattii infec-
tion of humans and animals, officials concede the public can do little to protect
themselves from infection (Knox, 2010). Additional research is needed to clarify
the epidemiology and drug susceptibilities of the various strains of C. gattii
present in the region to help inform treatment guidelines. Moreover, research-
ers need to learn more about the natural history and pathogenicity of the fungus
to further prevention, treatment, and intervention efforts (Datta et al., 2009a,b).
Heitman noted that research is ongoing to determine the nature of hypervirulence
(D’Souza et al., 2011); the differences between immune responses to C. gattii
and C. neoformans infections (Cheng et al., 2009); and why C. gattii can so read-
ily invade the cells of immunocompetent individuals (Kronstad et al., 2011; Ma
et al., 2009; Voelz and May, 2010).
FIGURE WO-27╇ Frogs in the Sierra Nevada region, being treated in baths containing a
fungicidal bacterium in hopes ofFigure
eliminating infection by the fungal pathogen (Bd) associ-
WO-27.eps
ated with the deadly disease: amphibian chytridiomycosis.
SOURCE: Photo by Anand Varma.
bitmap
89╛╛Fungicides, and more recently cutaneous bacteria of amphibians (e.g., J. lividum) known to
mechanisms by which these pathogens kill their hosts. Added to this is the unique
challenge of managing disease in hibernating animals in delicate underground
ecosystems. Biology of infectious diseases, however, is not part of the traditional
wildlife ecology education curriculum, Blehert remarked. Nor are speleologists,
tourists, recreational cavers, or hikers required to have such knowledge.
SCBI has used captive breeding to save several species from extinction. The
endangered black-footed ferret population was revived from just 18 individuals
in 1988 to a current population of 800 to 1,000 in the wild (Weidensaul, 2000).
Other animal species, including the golden lion tamarin, California condors,
Przewalski’s horses, and the scimitar-horned oryx, have also benefited from the
SCBI’s captive breeding program’s success.
Despite success with multiple species, establishing and maintaining captive
breeding programs is technically challenging. In the fall of 2010, SCBI devel-
oped a captive colony of the endangered Virginia big-eared bats (Corynorhinus
townsendii virginianus) in response to the threat posed by G. destructans. Spe-
cialist insect-eating bats, such as the Virginia big-eared bat, are notoriously dif-
ficult to keep in captivity. But, Padilla noted, in light of the possible extinction
of this endangered subspecies, SCBI decided to take on the “high risk” project
90╛╛Formerly known as the National Zoo’s Conservation and Research Center, the SCBI is an um-
brella organization for the Smithsonian’s global efforts to conserve species and train future genera-
tions of conservationists. See http://nationalzoo.si.edu/scbi/default.cfm.
91╛╛See contributed manuscript by Padilla in Appendix A (pages 296–312).
WORKSHOP OVERVIEW 81
92╛╛The Amphibian Ark carries out the ex situ components of the Amphibian Conservation Action
Plan developed by the World Conservation Union. For more information, see http://www.amphibian
ark.org/pdf/ACAP.pdf and www.amphibianark.org.
93╛╛See contributed manuscript by Weldon and Fisher in Appendix A (pages 355–367).
82 FUNGAL DISEASES
FIGURE WO-28╇ Panamanian golden frog (Atelopus zeteki). The Smithsonian National
Figureprogram
Zoo has established a captive breeding WO-28.eps
to help rescue this critically endangered
species. bitmap
SOURCE: Photo by Brian Gratwicke, Wikimedia commons.
of focus for preventing and managing all types of biological invasions, including
fungal pathogens, were discussed during the workshop as particularly promising:
Each of these approaches supports the overall goal of identifying and exploit-
ing common characteristics of invasive animals, plants, and microbes in order to
reduce their impact. To pursue this strategy requires “a new perspective, a new
thinking, a consideration of all alien introductions in a deliberate, truly compre-
hensive system,” ecologist Richard Mack has observed (Dybas, 2004, p. 618). “If
we do that, then we will have a sound science-based policy.”
84 FUNGAL DISEASES
———. 2011. The fungi: 1, 2, 3...5.1 million species? American Journal of Botany 98(3):
426–438.
Blackwell, M., D. S. Hibbett, J. W. Taylor, and J. W. Spatafora. 2006. Research coordination networks:
A phylogeny for kingdom Fungi (Deep Hypha). Mycologia 98:829–837.
Blackwell, M., R. Vigalys, R. James, Y. Timothy, and J.W. Taylor. 2009. Fungi. Eumycota: Mushrooms,
sac fungi, yeast, molds, rusts, smuts, etc. (Version 10). http://tolweb.org/Fungi/2377/2009.04.10
(accessed October 4, 2010).
Blanchette, R. A., A. M. Wilmering, and M. Baumeister. 1992. The use of green-stained wood caused
by the fungus Chlorociboria in intarsia masterpieces from the 15th century. Holzforschung
46:225–232.
Blehert, D. S., A. C. Hicks, M. Behr, C. U. Meteyer, B. M. Berlowski-Zier, E. L. Buckles, J. T. H.
Coleman, S. R. Darling, A. Gargas, R. Niver, J. C. Okoniewski, R. J. Rudd, and W. B. Stone.
2009. Bat white-nose syndrome: An emerging fungal pathogen? Science 323(5911):227.
Boyles, J. G., P. M. Cryan, G. F. McCraken, and H. Kunz. 2011. Conservation: Economic importance
of bats in agriculture. Science 332(6025):41–42.
Brasier, C. M. 2000. The rise of the hybrid fungi. Nature 405:134–135.
———. 2008. The biosecurity threat to the UK and global environment from international trade in
plants. Plant Pathology 57(5):792–808.
Brasier, C. M., and K. W. Buck. 2001. Rapid evolutionary changes in a globally invading fungal
pathogen (Dutch elm disease). Biological Invasions 3:223–233.
Brasier, C. M., and J. Webber. 2010. Sudden larch death. Nature 466:824–825.
Brasier, C. M., D. E. L. Cooke, and J. M. Duncan. 1999. Origin of a new Phytophthora pathogen
through interspecific hybridization. Proceedings of the National Academy of Sciences, USA 96:
5978–5983.
Briggs, C. J., A. R. Knapp, and V. T. Vredenburg. 2010. Enzootic and epizootic dynamics of the
chytrid fungal pathogen of amphibians. Proceedings of the National Academy of Sciences, USA
107 (21):9695–9700.
Bromenshenk, J. J., C. B. Henderson, C. H. Wick, M. F. Stanford, and A. W. Zulich. 2010. Iridovirus
and microsporidian linked to honey bee colony decline. PLoS ONE 5(10):e13181.
Brown, J. K. M., and M. S. Hovmøller. 2002. Aerial dispersal of pathogens on the global and conti-
nental scales and its impact on plant disease. Science 297(5581):537–541.
Brownstein, J. S., C. C. Freifeld, and L. C. Madoff. 2009. Digital disease detection—harnessing the
web for public health surveillance. New England Journal of Medicine 360(21):2153–2157.
Brucker, R. M., R. N. Harris, C. R. Schwantes, T. N. Gallaher, D. C. Flaherty, B. A. Lam, and K. P.
Minbiole. 2008. Amphibian chemical defense: Antifungal metabolites of the microsymbiont
Janthinobacterium lividum on the salamander Plethodon cinereus. Journal of Chemical Ecol-
ogy 34:1422–1429.
Buchen, L. 2010. Disease epidemic killing only U.S. bats. Nature 463(7278):144–145.
Buckley, M. 2008. The Fungal Kingdom: A report from the American Academy of Microbiology.
Washington, DC: American Academy of Microbiology.
Buckley, R. H., B. B. Wray, and E. Z. Belmaker. 1972. Extreme hyperimmunoglobulinemia E and
undue susceptibility to infection. Pediatrics 49(1):59–70.
Butterworth, M. H., M. A. Semenov, A. Barnes, D. Moran, J. S. West, and B. D. L. Fitt. 2010. North-
south divide: Contrasting impacts of climate change on crop yields in Scotland and England.
Journal of the Royal Society Interface 7:123–130.
Byrnes, E. J., R. J. Bildfell, S. A. Frank, T. G. Mitchell, K. A. Marr, and J. Heitman. 2009. Molecular
evidence that the range of the Vancouver Island outbreak of Cryptococcus gattii infection has
expanded into the Pacific Northwest in the United States. The Journal of Infectious Diseases
199:1081–1086.
Byrnes, E. J., W. Li, Y. Lewit, H. Ma, K. Voelz, P. Ren, D. A. Carter, V. Chaturvedi, R. J. Bildfell,
R. C. May, and J. Heitman. 2010. Emergence and pathogenicity of highly virulent Cryptococcus
gattii genotypes in the northwest United States. PLoS Pathogens 6(4):e1000850.
86 FUNGAL DISEASES
Capasso, L. 1998. 5300 years ago, the ice man used natural laxatives and antibiotics. The Lancet
352:1864.
Carlton, J. 2004. Invasions in the world’s oceans: How much do we know, and what does the future
hold? Presentations to the annual meeting of the American Institute of Biological Sciences,
2004. Available at: http://www.aibs.org/media-library
Casadevall, A. 2005. Fungal virulence, vertebrate endothermy, and dinosaur extinction: Is there a
connection? Fungal Genetics and Biology 42:98–106.
———. 2007. Determinants of virulence in the pathogenic fungi. Fungal Biology Reviews 21:130–132.
———. 2010. Emerging fungal pathogens—past, present, and future. Presentation given at the De-
cember 14–15, 2010, public workshop, “Fungal Diseases: An Emerging Challenge to Human,
Animal, and Plant Health,” Forum on Microbial Threats, Institute of Medicine, Washington,
D.C.
Casadevall, A., and L. A. Pirofski. 2003. The damage response framework of microbial pathogenesis.
Nature Reviews, Microbiology 1(1):17–24.
———. 2007. Accidental virulence, cryptic pathogenesis, martians, lost hosts, and the pathogenicity
of environmental microbes. Eukaryotic Cell 6:2169–2174.
Cash, R. A., and V. Narasimhan. 2000. Impediments to global surveillance of infectious diseases:
Consequences of open reporting in a global economy. Bulletin of the World Health Organiza-
tion 78(11):10.
Catenazzi, A., V. T. Vredenburg, and E. Lehr. 2010. Batrachochytrium dendrobatidis in the live frog
trade of Telmatobius (Anura: Ceratophyryidae) in the tropical Andes. Diseases of Aquatic
Organisms Preprint, 2010. http://web.me.com/vancevredenburg/Vances_site/Publications_files/
CatenazziVredenburgLehr2010.pdf.
CDC (Centers for Disease Control and Prevention). 2010. Emergence of Cryptococcus gattii—Pacific
Northwest, 2004–2010. Morbidity and Mortality Weekly Report 59(28):865–868.
Cendejas-Bueno, E. A., E. Gomez-Lopez, E. Mellado J. L. Rodriguez-Tudela, and M. Cuenca-
Estrella. 2010. Identification of pathogenic rare yeast species in clinical samples: Comparison
between phenotypical and molecular methods. Journal of Clinical Microbiology 48:1895–1899.
Chan, C. 2011. Valley fever cases likely to increase after Phoenix dust storm. The Arizona Republic.
July 18.
Chan, E. H., T. F. Brewer, L. C. Madoff, M. P. Pollack, A. L. Sonricker, M. Keller, C. C. Freifeld,
M. Blench, A. Mawudeku, and J. S. Brownstein. 2010. Global capacity for emerging infectious
disease detection. Proceedings of the National Academy of Sciences, USA 107:1–6.
Chaturvedi, V., D J. Springer, M. J. Behr, R. Ramani, X. Li, M. K. Peck, P. Ren, D. J. Bopp, B. Wood,
W. A. Samsonoff, C. M. Butchkoski, A. C. Hicks, W. B. Stone, R. J. Rudd, and S. Chaturvedi.
2010. Morphological and molecular characterizations of psychrophilic fungus Geomyces de-
structans from New York bats with white nose syndrome (WNS). PLoS ONE 5(5):e10783.
Chen, X. M. 2005. Epidemiology and control of stripe rust [Puccinia striiformis f. sp. tritici] on
wheat. Canadian Journal of Plant Pathology 27:314–337.
Cheng, P. Y., A. Sham, and J. W. Kronstadt. 2009. Cryptococcus gattii isolates from the British
Columbia Cryptococcosis outbreak induce less protective inflammation in a murine model of
infection than Cryptococcus neoformans. Infectious Immunity 77:4284–4294.
Chiller, T. M., J. N. Galgiani, and D. A. Stevens. 2003. Coccidioidomycosis. Infectious Disease Clin-
ics of North America 17:41–57.
Choffnes, E. R. 2008. Improving infectious disease surveillance. Bulletin of the Atomic Scientists
http://www.thebulletin.org/web-edition/op-eds/improving-infectious-disease-surveillance (ac-
cessed October 26, 2010).
Cliff, A., and P. Haggett. 2004. Time, travel and infection. British Medical Bulletin 69(1):87–99.
Cox, R. A., and D. M. Magee. 2004. Coccidioidomycosis: Host response vaccine development. Clini-
cal Microbiology Reviews 17:804–839.
Crum-Canflone, N. F. 2007. Coccidioidomycosis in the U.S. military: A review. Annals of the New
York Academy of Sciences 1111:112–121.
WORKSHOP OVERVIEW 87
Cryan, P. M., C. Uphoff Meteyer, J. G. Boyles, and D. S. Blehert. 2010. Wing pathology of white-
nose syndrome in bats suggests life-threatening disruption of physiology. BMC Biology 8(135).
Cunningham, A. A., and P. Daszak. 1998. Extinction of a species of land snail due to infection with
a microsporidian parasite. Conservation Biology 12:1523–1739.
Currie, C. R., U. G. Mueller, and D. Malloch. 1999. The agricultural pathology of ant fungus gardens.
Proceedings of the National Academy of Sciences, USA 96(14):7998–8002.
Currie, C. R., B. Wong, A. E. Stuart, T. R. Schultz, S. A. Rehner, U. G. Mueller, G. H. Sung, J. W.
Spatafora, and N. A. Straus. 2003. Ancient tripartite coevolution in the attine ant-microbe sym-
biosis. Science 299:386–388.
Cutler, J. E., S. G. Deepe, Jr., and B. S. Klein. 2007. Advances in combating fungal diseases: Vaccines
on the threshold. Nature Reviews Microbiology 5:13–28.
Dadachova, E., and A. Casadevall. 2008. Ionizing radiation: How fungi cope, adapt, and exploit with
the help of melanin. Current Opinions in Microbiology 11(6):525–531.
Dadachova, E., R. A. Bryan, X. Huang, T. Moadel, A. D. Schweitzer, P. Aisen, J. D. Nosanchuk, A.
Casadevall. 2007. Ionizing radiation changes the electronic properties of melanin and enhances
the growth of melanized fungi. PLoS ONE 2(5):e457.
Daszak, P. 2010. Global capacity for coordinated surveillance, detection, and response to emerging
diseases of wildlife. Presentation given at the December 14–15, 2010, public workshop, “Fungal
Diseases: An Emerging Challenge to Human, Animal, and Plant Health,” Forum on Microbial
Threats, Institute of Medicine, Washington D.C.
Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2000. Emerging infectious diseases of wildlife—
threats to biodiversity and human health. Science 287(5452):443–449.
———. 2003. Infectious disease and amphibian population declines. Diversity and Distributions
9:141–150.
Daszak, P., A. Strieby, A. A. Cunningham, J. E. Longcore, C. C. Brown, and D. Porter. 2004. Experi-
mental evidence that the bullfrog (Rana catesbeiana) is a potential carrier of chytridiomycosis,
an emerging fungal disease of amphibians. Herpetological Journal 14:201–207.
Datta, K., K. H. Bartlett, R. Baer, E. Byrnes, E. Galanis, J. Heitman, L. Hoang, M. J. Leslie, L.
MacDougall, S. S. Magill, M. G. Morshed, and K. A. Marr. 2009a. Spread of Cryptococ-
cus gattii: Into Pacific Northwest region of the United States. Emerging Infectious Diseases
15(8):1185–1191.
Datta, K., K. H. Bartlett, and K. A. Marr. 2009b. Cryptococcus gattii: Emergence in western North
America: Exploitation of a novel ecological niche. Interdisciplinary Perspectives on Infectious
Diseases. Article ID 176532, 8 pages doi:10.1155/2009/176532.
Daughtrey, M. L., C. R. Hibben, K. O. Britton, M. T. Windham, and S. C. Redlin. 1996. Dogwood
anthracnose: Understanding a disease new to North America. Plant Disease 80(4):349–358.
Davis, S. D., J. Schaller, and R. J. Wedgwood. 1996. Job’s syndrome: Recurrent, “cold,” staphylococ-
cal abscesses. The Lancet 1(7445):1013–1015.
De Lucca, A. J. 2007. Harmful fungi in both agriculture and medicine. Revista iberoamericana de
micología 24:11.
Dentinger, B. T. M., D. J. Lodge, A. B. Munkacsi, D. E. Desjardin, and D. J. McLaughlin. 2009. Phy-
logenetic placement of an unusual coral mushroom challenges the classic hypothesis of strict co-
evolution in the Apterostigma pilosium group ant-fungus mutualism. Evolution 63:2172–2178.
Desjardin, D. E., B. A. Perry, D. J. Lodge, C. V. Stevani, and E. Nagasawa. 2010. Luminescent my-
cena: New and noteworthy species. Mycologia 102(2):459–477.
Desprez-Loustau, M. L., C. Robin, M. Buee, R. Courtecuisse, J. Garbaye, F. Suffert, I. Sache, and
D. M. Rizzo. 2007. The fungal dimension of biological invasions. Trends in Ecology and Evolu-
tion 22(9):472–480.
Dixon, D. M., M. M. McNeil, M. L. Cohen, B. G. Gellin, and J. R. La Montagne. 1996. Fungal infec-
tions: A growing threat. Public Health Reports 111(3):226–235.
Drew, A., E. J. Allen, and L. J. Allen. 2006. Analysis of climatic and geographic factors affecting the
presence of chyridiomycosis in Australia. Diseases of Aquatic Organisms 68:245–250.
88 FUNGAL DISEASES
Garbelotto, M., D. M. Rizzo, and L. Marais. 2002. Phytophthora ramorum and sudden oak death
in California. Chemical control. In: Proceedings of the 5th Symposium on California Oak
Woodlands, edited by R. Standiford and D. McCreary. U.S. Department of Agriculture, Forest
Service, pp. 811–818.
Garcia-Solache, M. A., and A. Casadevall. 2010. Global warming will bring new fungal diseases for
mammals. mBio 1:e00061–10.
Gargas, A. M., T. Trest, M. Christensen, T. J. Volk, and D. S. Blehert. 2009. Geomyces destructans
sp. nov. associated with bat white-nose syndrome. Mycotaxon 108:147–154.
Garner, C. D., J. K. Starr, P. L. McDonough, and C. Altier. 2010. Molecular identification of veterinary
yeast isolates by use of sequence-based analysis of the D1/D2 region of the large ribosomal
subunite. Journal of Clinical Microbiology 48:2140–2146.
Garrett, K. A., S. P. Dendy, E. E. Frank, M. N. Rouse, and S. E. Travers. 2006. Climate change ef-
fects on plant disease: Genomes to ecosystems. Annual Reviews in Phytopathology 44:489–509.
Giraud, T., P. Gladieux, and S. Gavrilets. 2010. Linking the emergence of fungal plant diseases with
ecological speciation. Trends in Ecology and Evolution 30:101–109.
Goddard, M. R., H. C. Godfray, and J. A. Burt. 2005. Sex increases the efficacy of natural selection
in experimental yeast populations. Nature. 434:636–640.
Goldman, D. L., H. Khine, J. Abadi, D. J. Lindenberg, L. Pirofski, R. Niang, and A. Casadevall.
2001. Serological evidence for Cryptococcus neoformans infection in early childhood. Pedi-
atrics 107:e66.
Goldman, D. L., J. Davis, F. Bommarito, X. Shao, and A. Casadevall. 2006. Enhanced allergic in-
flammation and airway responsiveness in rats with chronic Cryptococcus neoformans infection:
Potential role for fungal pulmonary infection in the pathogenesis of asthma. Journal of Infec-
tious Diseases 193:1178–1186.
Goss, E. M., M. Larsen, G. A. Chastagner, D. R. Givens, and N. J. Grünwald. 2009. Population
genetic analysis infers migration pathways of Phytophthora ramorum in U.S. nurseries. PLoS
Pathogenics 5(9):e1000583.
Gostinčar, C., M. Grube, S. de Hoog, P. Zalar, and N. Gunde-Cimerman. 2010. Extremotolerance in
fungi: Evolution on the edge. FEMS Microbiology Ecology 71:2–11.
Greer, A., N. Victoria, and D. Fisman. 2008. Climate change and infectious diseases in North
America: The road ahead. Canadian Medical Association Journal 178:6.
Grünwald, N. J., E. M. Gross, and C. M. Press. 2008. Phytophthora ramorum: A pathogen with a
remarkably wide host range causing sudden oak death on oaks and ramorum blight on woody
ornamentals. Molecular Plant Pathology 9(5):1–11.
Hannukkala, A., O. Kaukoranta, T. Lehtinen, and A. Rahkonen. 2007. Late-blight epidemics on potato
in Finland, 1933–2002: Increased and earlier occurrence of epidemics associated with climate
change and lack of rotation. Plant Pathology 56:167–176.
Hardman, R. 2011. Britain’s forests: 10k acres of trees cut down to stop pathogen. Dailymail, Janu-
ary 27.
Harris, R. N., R. M. Brucker, J. B. Walke, M. H. Becker, C. R. Schwantes, D. C. Flaherty, B. A. Lam,
D. C. Woodhams, C. J. Briggs, V. T. Vredenburg, and K. P. C. Minbiole. 2009. Skin microbes
on frogs prevent morbidity and mortality caused by a lethal skin fungus. The ISME Journal
2009:1–7.
Harvell, C. D., C. E. Mitchell, J. R. Ward, S. Altizer, A. P. Dobson, R. S. Ostfeld, and M. D.
Samuel. 2002. Climate warming and disease risks for terrestrial and marine biota. Science
296(5576):2158–2162.
Hawksworth, D. L. 1991. The fungal dimension of biodiversity: Magnitude, significance, and conser-
vation. Mycology Research 6:641–655.
———. 2001. The magnitude of fungal diversity: The 1.5 million species estimate revisited. Myco-
logical Research 105(12):1422–1432.
Hector, R. F., and R. Laniado-Laborin. 2005. Coccidioidomycosis—a fungal disease of the Americas.
PLoS Medicine 2:0015–0018.
90 FUNGAL DISEASES
Hector, R. F., B. L. Zimmer, and D. Pappagianis. 1990. Evaluation of nikkomycins X and Z in murine
models of coccidioidomycosis, histoplasmosis, and blastomycosis. Antimicrobial Agents and
Chemotherapy 34:587–593.
Heitman, J. 2006. Sexual reproduction and the evolution of microbial pathogens. Current Biology
16:R711–R725.
———. 2009. Love the one you’re with. Nature 460(13):807–808.
Heymann, D. L., and G. Rodier. 2004. Global surveillance, national surveillance, and SARS. Emerg-
ing Infectious Diseases 10(2):3.
Hibbett, D. M., M. Binder, J. F. Bischoff, M. Blackwell, P. F. Cannon, O. Eriksson, S. Huhndorf, T. Y.
James, P. M. Kirk, R. Lücking, T. Lumbsch, F. Lutzoni, P. B. Matheny, D. J. McLaughlin, M. J.
Powell, S. Redhead, C. L. Schoch, J. W. Spatafora, J. A. Stalpers, R. Vilgalys, M. C. Aime, A.
Aptroot, R. Bauer, D. Begerow, G. L. Benny, L. A. Castlebury, P. W. Crous, Y.-C. Dai, W. Gams,
D. M. Geiser, G. W. Griffith, D. L. Hawksworth, V. Hofstetter, K. Hosaka, R. A. Humber, K.
Hyde, U. Kõljalg, C. P. Kurtzman, K.-H. Larsson, R. Lichtwardt, J. Longcore, A. Miller, J.-M.
Moncalvo, S. Mozley Standridge, F. Oberwinkler, E. Parmasto, J. D. Rogers, L. Ryvarden, J. P.
Sampaio, A. Schuessler, J. Sugiyama, J. W. Taylor, R. G. Thorn, L. Tibell, W. A. Untereiner,
C. Walker, Z. Wang, A. Weir, M. Weiss, M. White, K. Winka, Y.-J. Yao, and N. Zhang. 2007.
A higher-level phylogenetic classification of the Fungi. Mycological Research 111: 509–547.
Holland, S. M. 2010. Chronic granulomatous disease. Clinical Reviews in Allergy and Immunology
38(1):3–10.
Holland, S. M., and D. C. Vinh. 2009. Yeast infections—human genetics on the rise. New England
Journal of Medicine 361:1798–1801.
Holland, S. M., F. R. DeLeo, H. Z. Elloumi, A. P. Hsu, G. Uzel, N. Brodsky, A. F. Freeman, A.
Demidowich, J. Davis, M. L. Turner, V. L. Anderson, D. N. Darnell, P. A. Welch, D. B. Kuhns,
D. M. Frucht, H. L. Malech, J. I. Gallin, S. D. Kobayashi, A. R. Whitney, J. M. Voyich, J. M.
Musser, C. Woellner, A. A. Schäffer, J. M. Puck, and B. Grimbacher. 2007. STAT3 mutations in
the hyper-IgE syndrome. New England Journal of Medicine 357:1608–1619.
Hooper, R. G., M. R. Lennartz, and D. H. Muse. 1991. Heart rot and cavity tree selection by red-
cockaded woodpeckers. Journal of Wildlife Management 55(2):323–327.
Hovmøller, M. 2010. Rapid global spread of aggressive strains of Puccinia striiformis on wheat—
origins, causes, and consequences. Presentation given at the December 14–15, 2010, public
workshop, “Fungal Diseases: An Emerging Challenge to Human, Animal, and Plant Health,”
Forum on Microbial Threats, Institute of Medicine, Washington, DC.
Hovmøller, M. S., A. H. Yahyaoui, and E. A. Milus. 2008. Rapid global spread of two aggressive
strains of a wheat rust fungus. Molecular Ecology 17:3818–3826.
Hovmøller, M. S., S. Walter, and A. F. Justesen. 2010. Escalating threat of wheat rusts. Science
329:369.
Hufnagel, L., D. Brockmann, and T. Geisel. 2004. Forecast and control of epidemics in a globalized
world. Proceedings of the National Academy of Sciences, USA 101(42):15124–15129.
IOM (Institute of Medicine) 1992. Emerging infections. Washington, DC: National Academy Press.
———. 2003. Microbial threats to heath. Washington, DC: The National Academies Press.
———. 2007. Global infectious disease surveillance and detection: Assessing the challenges. Work-
shop summary. Washington, DC: The National Academies Press.
———. 2008a. Global climate change and extreme weather events: Understanding the contribu-
tions to infectious disease emergence: Workshop summary. Washington, DC: The National
Academies Press.
———. 2008b. Vector borne diseases: Understanding the environmental, human health, and ecologi-
cal connections. Washington, DC: The National Academies Press.
———. 2009. Microbial evolution and co-adaptation: A tribute to the life and scientific legacies
of Joshua Lederberg. Workshop summary. Washington, DC: The National Academies Press.
———. 2010. Infectious disease movement in a borderless world. Washington, DC: The National
Academies Press.
WORKSHOP OVERVIEW 91
Iqbal, N., E. E. DeBess, R. Wohrle, B. Sun, R. J. Nett, A. M. Ahlquist, T. Chiller, and S. R. Lockhart.
2010. Correlation of genotype and in vitro susceptibilities of Cryptococcus gattii strains from
the Pacific Northwest of the United States. Journal of Clinical Microbiology 48(2):539–544.
IUCN (The World Conservation Union). 2005. Amphibian Conservation Action Plan. http://www.
amphibianark.org/pdf/ACAP.pdf (accessed November 20, 2010).
James, T. Y., K. Kauff, C. L. Schoch, P. B. Matheny, V. Hofstetter, C. J. Cox, G. Celio, C. Gueidan,
E. Fraker, J. Miadlikowska, H. T. Lumbsch, A. Rauhut, V. Reeb, A. E. Arnold, A. Amtoft, J. E.
Stajich, K. Hosaka, G. H. Sung, D. Johnson, B. O’Rourke, M. Crockett, M. Binder, J. M. Curtis,
J. C. Slot, Z. Wang, A. W. Wilson, A. Schüssler, J. E. Longcore, K. O’Donnell, S. Mozley-
Standridge, D. Porter, P. M. Letcher, M. J. Powell, J. W. Taylor, M. M. White, G. W. Griffith,
D. R. Davies, R. A. Humber, J. B. Morton, J. Sugiyama, A. Y. Rossman, J. D. Rogers, D. H.
Pfister, D. Hewitt, K. Hansen, S. Hambleton, R. A. Shoemaker, J. Kohlmeyer, B. Volkmann-
Kohlmeyer, R. A. Spotts, M. Serdani, P. W. Crous, K. W. Hughes, K. Matsuura, E. Langer,
G. Langer, W. A. Untereiner, R. Lücking, B. Büdel, D. M. Geiser, A. Aptroot, P. Diederich,
I. Schmitt, M. Schultz, R. Yahr, D. S. Hibbett, F. Lutzoni, D. J. McLaughlin, J. W. Spatafora,
and R. Vilgalys. 2006. Restructuring the early evolution of Fungi using a six gene phylogeny.
Nature 443:818–822.
James, T. Y., A. P. Litvintseva, R. Vilgalys, J. A. Morgan, J. W. Taylor, M. C. Fisher, L. Berger, C.
Weldon, L. du Preez, and J. E. Longcore. 2009. Rapid global expansion of the fungal dis-
ease chytridiomycosis into declining and healthy amphibian populations. PLoS Pathogens
5(5):e1000458 1–12.
Jebara, K. B. 2004. Surveillance, detection and response: Managing emerging diseases at national and
international levels. OIE Revue Scientifique et Technique 23(2):709–715.
Jeger, M. J., and M. Pautasso. 2008. Plant disease and global change—the importance of long-term
data sets. New Phytologist 177:8–11.
Jones, K. E., N. G. Patel, M. A. Levy, A. Storeygard, D. Balk, J. L. Gittleman, and P. Daszak. 2008.
Global trends in emerging infectious diseases. Nature 451:990–993.
Jones, M. D. M., I. Forn, C. Gadelha, M. J. Egan, D. Bass, R. Massana, T. A. Richards. 2011. Discov-
ery of novel intermediate forms redefines the fungal tree of life. Nature Published online May
11, 2011. doi:10.1038/nature09984.
Judelson, H. S., and F. A. Blanco. 2005. The spores of phytophthora: Weapons of the plant destroyer.
Nature Reviews/Microbiology 3:47–58.
Jumpponen, A., and K. L. Jones. 2009. Massively parallel 454 sequencing indicates hyperdiverse fun-
gal communities in temperate Quercus macrocarpa phyllosphere. New Phytologist 184:438–448.
Karesh, W. B., R. A. Cook, E. L. Bennet, and J. Newcomb. 2005. Wildlife trade and global disease
emergence. In Emerging Infectious Diseases.
Kauserud, H., L. C. Stige, J. O. Vik, R. H. Økland, K. Høiland, N. C. Stenseth. 2008. Mush-
room fruiting and climate change. Proceedings of the National Academy of Sciences, USA
105(10):3811–3814.
Keller, N. P., G. Turner, and J. W. Bennett. 2005. Fungal secondary metabolism from biochemistry to
genomics. Nature Reviews/Microbiology 3:937–947.
Keller, R., and C. Perrings. 2010. International policy options to reduce the harmful impacts of alien
invasive species. UNEP Ecosystem Services Economics Working Papers, Nairobi, UNEP
Kelly, M., D. Shaari, Q. Guo, and D. Liu. Spatial modeling of sudden oak death nationwide. 2005.
U.S. Department of Agriculture, Forest Service Gen. Tech. Rep. PSW-GTR-196-006-063. http://
www.suddenoakdeath.org/pdf/KellyetalSOD2-22-05.pdf (accessed May 3, 2011).
Kidd, S. E., F. Hagen, R. L. Tscharke, M. Huynh, K. H. Bartlett, M. Fyfe, L. MacDougall, T.
Boekhout, K. J. Kwon-Chung, and W. Meyer. 2004. A rare genotype of Cryptococcus gattii
caused the cryptococcosis outbreak on Vancouver Island (British Columbia, Canada). Proceed-
ings of the National Academy of Sciences, USA 101(49):17258–17263.
92 FUNGAL DISEASES
Lonsdale, D., and J. N. Gibbs. 1996. Effects of climate change on fungal diseases of trees. In: Fungi
and environmental change, edited by J. E. Frankland, N. Magan, and G. M. Gadd. British
Mycological Society, Symp vol. XX. Cambridge, UK: Cambridge University Press, Pp. 1–19.
Loo, J. 2009. Ecological impacts of non-indigenous invasive fungi as forest pathogens. Biological
Invasions 11(1):81–96.
Lötters, S., J. Kielgast, J. Bielby, S. Schmidtlein, J. Bosch, M. Veith, S. F. Walker, M. C. Fisher, and
D. Rödder. 2010. The link between rapid enigmatic amphibian decline and the globally emerg-
ing chytrid fungus. EcoHealth 1–15.
Loustau, D. 2006. Climate change impacts on extensively managed forest: a modelling approach,
Wilton Park Conference. See: http://www.forestry.gov.uk/forestry/INFD-6VKDVB (accessed
June 22, 2011).
Lutzoni, F., M. Pagel, and V. Reeb. 2001. Major fungal lineages are derived from lichen symbiotic
ancestors. Nature 411:937–940.
Lutzoni, F., F. Kauff, C. J. Cox, D. McLaughlin, G. Celio, B. Dentinger, M. Padamsee, D. Hibbett, T. Y.
James, E. Baloch, M. Grube, V. Reeb, V. Hofstetter, C. Schoch, A. E. Arnold, J. Miadlikowska,
J. Spatafora, D. Johnson, S. Hambleton, M. Crockett, R. Shoemaker, G. H. Sung, R. Lucking, T.
Lumbsch, K. O’Donnell, M. Binder, P. Diederich, D. Ertz, C. Gueidan, K. Hansen, R. C. Harris,
K. Hosaka, Y. W. Lim, B. Matheny, H. Nishida, D. Pfister, J. Rogers, A. Rossman, I. Schmitt,
H. Sipman, J. Stone, J. Sugiyama, R. Yahr, and R. Vilgalys. 2004. Assembling the fungal tree
of life: Progress, classification, and evolution of subcellular traits. American Journal of Botany
91(10):1446–1480.
Ma, H., F. Hagen, D. J. Stekel, S. A. Johnston, E. Sionov, R. Falk, I. Polacheck, T. Boekhout, and
R. C. May. 2009. The fatal fungal outbreak on Vancouver Island is characterized by enhanced
intracellular parasitism driven by mitochondrial regulation. Proceedings of the National Acad-
emy Sciences, USA 106(31):12980–12985.
MacDougall, L., S. E. Kidd, E. Galanis, S. Mak, M. J. Leslie, P. R. Cieslak, J. W. Kronstad, M. G.
Morshed, and K. H. Bartlett. 2007. Spread of Cryptococcus gattii in British Columbia, Canada,
and detection in the Pacific Northwest, USA. Emerging Infectious Diseases 13(1):42–50.
MacLeod, A. M. Pautasso, M. J. Jeger, and R. Haines-Young. 2010. Evolution of the international
regulation of plant pests and challenges for future plant health. Food Security 2:49–70.
Madoff, L. C. 2004. ProMED-mail: An early warning system for emerging diseases. Clinical Infec-
tious Diseases 39(2):227–232.
Mak, S., B. Klinkenberg, K. Bartlett, and M. Fyfe. 2010. Ecological niche modeling of Cryptococcus
gattii in British Columbia, Canada. Environmental Health Perspectives 118:653–658.
Márquez, L. M., R. S. Redman, R. J. Rodriguez, and M. J. Roossinck. 2007. A virus in a fungus in a
plant: Three-way symbiosis required for thermal tolerance. Science 315:513–515.
Mascheretti, S. P., J. P. Croucher, A. Vettraino, S. Prospero, and M. Garbelotto. 2008. Reconstruction
of the sudden oak death epidemic in California through microsatellite analysis of the pathogen
Phytophthora ramorum. Molecular Ecology 17:2755–2768.
Mascheretti, S. P., J. P. Croucher, M. Kozanitas, L. Baker, and M. Garbelotto. 2009. Genetic epide-
miology of the sudden oak death pathogen Phytophthora ramorum in California. Molecular
Ecology 18(22):4577–4590.
McCullough, D. G., T. T. Work, J. F. Cavey, A. M. Liebhold, and D. Marshall. 2006. Interceptions
of nonindigenous plant pests at U.S. ports of entry and border crossings over a 17-year period.
Biological Invasions 8:611–630.
McLaughlin, D. J., D. S. Hibbett, F. Lutzoni, J. W. Spatafora, and R. Vilgalys. 2009. The search for
the fungal tree of life. Trends in Microbiology 17(11):488–497.
Meentemeyer, R., D. Rizzo, W. Mark, and E. Lotz. 2004. Mapping the risk of establishment and
spread of sudden oak death in California. Forest Ecology and Management 200(1–3):195–214.
Meteyer, C. U., E. L. Buckles, D. S. Blehert, A. C. Hicks, D. E. Green, V. Shearn-Bochsler, N. J.
Thomas, A. Gargas, and M. J. Behr. 2009. Histopathologic criteria to confirm white-nose syn-
drome in bats. Journal of Veterinary Diagnostic Evaluation 21(4):411–414.
94 FUNGAL DISEASES
Miller, O. K., Jr., T. Henkel, T. Y. James, and S. L. Miller. 2001. Pseudotulostoma, a remarkable new
volvategenus in the Elaphomycetaceae from Guyana. Mycological Research 105:1268–1272.
Milus, E. A., E. Seyran, and R. McNew. 2006. Aggressiveness of Puccinia striiformis f. sp. tritici
isolates in the South-Central United States. Plant Disease 90:847–852.
Milus, E. A., K. Kristensen, and M. S. Hovmøller. 2009. Evidence for increased aggressiveness in a
recent widespread strain of Puccinia striiformis f. sp. tritici causing stripe rust wheat. Phyto-
pathology 99:89–94.
Money, N. P. 2007. The triumph of the fungi: A rotten history. New York: Oxford University Press.
Morgan, J. A. T., V. T. Vredenburg, L. J. Rachowicz, R. A. Knapp, M. J. Stice, T. Tunstall, R. E.
Bingham, J. M. Parker, J. E. Longcore, C. Moritz, C. J. Briggs, and J. W. Taylor. 2007. Popula-
tion genetics of the frog-killing fungus Batrachochytrium dendrobatidis. Proceedings of the
National Academy of Sciences, USA 104:13845–13850.
Morse, S. 2004. Emerging infections: Microbial invaders discover new territory. Presentation to the
annual meeting of the American Institute of Biological Sciences, 2004. http://www.aibs.org/
media-library/ (accessed June 22, 2011).
Morse, S., S. B. Hatch, Rosenberg, and J. Woodall. 1996. Global monitoring of emerging diseases:
Design for a demonstration program. Health Policy 38:135–153.
Munkacsi, A. B., J. J. Pan, P. Villesen, U. G. Mueller, M. Blackwell, and D. J. McLaughlin. 2004.
Convergent coevolution in the domestication of coral mushrooms by fungus-growing ants.
Proceedings of the Royal Society of London, B 271:1777–1782.
Nardi, J. B., C. M. Bee, L. A. Miller, N. H. Nguyen, S.-O. Suh, and M. Blackwell. 2006. Communi-
ties of microbes that inhabit the changing hindgut landscape of a subsocial beetles. Arthropod
Structure & Development 35:57–68.
Nucci, M., and K. A. Marr. 2005. Emerging fungal diseases. Clinical Infectious Diseases 41(4):
521–526.
Nürnberger, T., F. Brunner, B. Kemmerling, and L. Piater. 2004. Innate immunity in plants and ani-
mals: Striking similarities and obvious differences. Immunology Reviews 198:249–266.
O’Donnell, A. 2006. Invasive species: More aggressive import screening is cost-effective, says study.
Land Letter: Natural Resources Weekly Report.
Oerke, E. C., H. W. Dehne, F. Schohnbeck, and A. Weber. 1995. Crop production and crop protec-
tion: Estimated losses in major food and cash crops. Amsterdam, The Netherlands and New
York: Elsevier.
Oklahoma Department of Wildlife Conservation and U.S. Fish and Wildlife Service. 2010. Bat
fungus documented in Oklahoma, www.wildlifedepartment.com/newsreleasearchive/05-10nr.
htm#Bat_fungus_documented_in_Oklahoma (accessed May 4, 2011).
Ortoneda, M., J. Guarro, M. P. Madrid, Z. Caracuel, M. I. Roncero, E. Mayayo, and A. Di Pietro.
2004. Fusarium oxysporum as a multihost model for the genetic dissection of fungal virulence
in plants and mammals. Infection and Immunity 72:1760–1766.
Ostrosky-Zeichner, L., A. Casadevall, J. N. Galgiani, F. C. Odds, and J. H. Rex. 2010. An insight into
the antifungal pipeline: Selected new molecules and beyond. Nature Reviews Drug Discovery
9(9):719–727.
Park, B. J., K. A. Wannemuehler, B. J. Marston, N. Govender, P. G. Pappas, and T. M. Chiller. 2009.
Estimation of the current global burden of cryptococcal meningitis among persons living with
HIV/AIDS. AIDS 23:525–530.
Parke, J. L., and S. Lucas. 2008. Sudden oak death and ramorum blight. The Plant Health Instructor.
doi:10.1094/PH-I-2008-0227-01.
Pautasso, M., K. Dehnen-Schmutz, O. Holdenrieder, S. Pietravalle, N. Salama, M. Jeger, E. Lange,
and S. Hehl-Lange. 2010. Plant health and global change—some implications for landscape
management. Biological Reviews 85(4):729–755.
Perrings, C., S. Burgiel, M. Lonsdale, H. Mooney, and M. Williamson. 2010. International coopera-
tion in the solution to trade-related invasive species risks. Annals of the New York Academy of
Sciences 1195:198–212.
WORKSHOP OVERVIEW 95
Pfaller, M. A., and D. J. Diekema. 2010. Epidemiology of invasive mycoses in North America. Criti-
cal Reviews in Microbiology 36(1):1–53.
Pimentel, D., R. Zuniga, and D. Morrison. 2005. Update on the environmental and economic costs as-
sociated with alien-invasive species in the United States. Ecological Economics 52(3):273–288.
Pirofski, L. A., and A. Casadevall. 2008. The damage-response framework of microbial pathogenesis
and infectious diseases. Experimental Biology and Medicine 635:135–146.
Platt, J. 2010. Bad news for bats: Deadly white-nose syndrome still spreading. http://www.scientifi-
camerican.com/blog/post.cfm?id=bad-news-for-bats-deadly-white-nose-2010-02-20 (accessed
May 12, 2010).
Porter, T. M., C. W. Schadt, L. Rizvi, A. P. Martin, S. K. Schmidt, L. Scott-Denton, R. Vilgalys, and
J. M. Moncalvo. 2008. Widespread occurrence and phylogenetic placement of a soil clone group
adds a prominent new branch to the fungal tree of life. Molecular Phylogenetics and Evolution
46:635–664.
Pounds, J. A., M. R. Bustamante, L. A. Coloma, J. A. Consuegra, M. P. L. Fogden, P. N. Foster, E.
La Marca, K. L. Masters, A. Merino-Viteri, R. Puschendorf, S. R. Ron, G. A. Sánchez-Azofeifa,
C. J. Still, and B. E. Young. 2006. Widespread amphibian extinctions from epidemic disease
driven by global warming. Nature 439(7073):161–167.
Puechmaille, S. J., P. Verdeyroux, H. Fuller, M. Ar Gouilh, M. Bekaert, and E. C. Teeling. 2010.
Whitenose syndrome fungus (Geomyces destructans) in bat, France. Emerging Infectious Dis-
eases 16(2):290–293.
Puechmaille, S. J., G. Wibbelt, V. Korn, H. Fuller, F. Forget, K. Muhldorfer, A. Kurth, B. Wieslaw,
C. Borel, T. Bosch, T. Cherezy, M. Drebet, T. Gorfol, A. J. Haarsma, F. Herhaus, G. Hallart, M.
Hammer, C. Jungmann, Y. Le Bris, L. Lutsar, M. Masing, B. Mulkens, K. Passior, M. Starrach,
M. Wojtaszewski, U. Zophel, and E. C. Teeling. 2011. Pan-European distribution of white-nose
syndrome fungus (Geomyces destructans) not associated with mass mortality. PLoS Pathogens
6(4):e19167.
Qaummen, D. 2010. Bat crash. National Geographic Magazine, December, Pp. 126–137.
Rachowicz, L. J., J. M. Hero, R. A. Alford, J. W. Taylor, and J. A. T. Morgan. 2005. The novel and
endemic pathogen hypotheses: Competing explanations for the origin of emerging infectious
diseases of wildlife. Conservation Biology 19:1441–1448.
Rizzo, D. M. 2005. Exotic species and fungi: Interactions with fungal, plant and animal communi-
ties. In: The fungal community, 3rd ed., edited by J. Dighton, P. Oudemans, and J. White. CRC
Press, Pp. 857–877.
———. 2010. Emergence of Phytophthora ramorum in Europe and North America. Presentation
given at the December 14–15, 2010, public workshop, “Fungal Diseases: An Emerging Chal-
lenge to Human, Animal, and Plant Health,” Forum on Microbial Threats, Institute of Medicine,
Washington, DC.
Rizzo, D. M., and M. Garbelotto. 2003. Sudden oak death: Endangering California and Oregon forest
ecosystems. Frontiers in Ecology and the Environment 1(5):197–204.
Rizzo, D. M., M. Garbelotto, and E. M. Hansen. 2005. Phytophthora ramorum: Integrative research
and management of an emerging pathogen in California and Oregon forests. Annual Review of
Phytopathology 43(1):309–335.
Roach, J. 2011. Caterpillar fungus making Tibetan herders rich. National Geographic News. http://
news.nationalgeographic.com/news/2011/04/110427-fungus-caterpillars-tibet-china-herders-
science (accessed June 22, 2011).
Robert, V. A., and A. Casadevall. 2009. Vertebrate endothermy restricts most fungi as potential patho-
gens. Journal of Infectious Disease 200:1623–1626.
Rödder, D., J. Kielgast, J. Bielby, J. Bosch, T. J. W. Garner, S. Schmidtlein, M. Veith, S. Walker, M.€C.
Fisher, and S. Lötters. Global amphibian extinction risk assessment for the panzootic chytrid
fungus. Diversity 1:52–66.
Rodriguez, R. J., J. F. White, Jr., A. E. Arnold, and R. S. Redman. 2009. Fungal endophytes: Diversity
and functional roles. New Phytologist 182:314–330.
96 FUNGAL DISEASES
Sexton, A. C., and B. J. Howlett. 2006. Parallels in fungal pathogenesis on plant and animal hosts.
Eukaryotic Cell 5:1941–1949.
Sharpton, T. J., J. E. Stajich, S. D. Rounsley, M. J. Gardner, J. R. Wortman, V. S. Jordar, R. Maiti, C. D.
Kodira, D. E. Neafsey, Q. D. Zeng, C. Y. Hung, C. McMahan, A. Muszewska, M. Grynberg,
M. A. Mandel, E. M. Kellner, B. M. Barker, J. N. Galgiani, M. J. Orbach, T. N. Kirkland, G.€T.
Cole, M. R. Henn, B. W. Birren, and J. W. Taylor. 2009. Comparative genomic analyses of the
human fungal pathogens Coccidioides and their relatives. Genome Research 19:1722–1731.
Shoham, S., and S. M. Levitz. 2005. The immune response to fungal infections. British Journal of
Haematology 129:569–582.
Siddiqui, S., V. L. Anderson, D. M. Hilligoss, M. Abinun, T. W. Kuijpers, H. Masur, F. G. Witebsky,
Y. R. Shea, J. I. Gallin, H. L. Malech, and S. M. Holland. 2007. Fulminant mulch pneumonitis:
An emergency presentation of chronic granulomatous disease. Clinical Infectious Diseases
45:673–681.
Skerratt, L. F., L. Berger, R. Speare, S. Cashins, K. R. McDonald, A. D. Phillott, H. B. Hines, and N.
Kenyon. 2007. Spread of chytridiomycosis has caused the rapid global decline and extinction
of frogs. EcoHealth 4:125–134.
Smith, K., F. M. Behrens, L. M. Schloegel, N. Marano, S. Burgiel, and P. Daszak. 2009. Reducing
the risks of the wildlife trade. Science 324(5927):594–595.
Snieszko, S. F. 1974. The effects of environmental stress on outbreaks of infectious diseases of fishes.
Journal of Fish Biology 6(2):197–208.
Sorrell, T. C. 2001. Cryptococcus neoformans variety gattii. Medical Mycology 39(2):155–168.
Sprague, S. J., S. J. Marcroft, H. L. Hayden, and B. J. Howlett. 2006. Major gene resistance to black-
leg in Brassica napus overcome within three years of commercial production in southeastern
Australia. Plant Disease 90:190–198.
Stajich, J. E., M. L. Berbee, M. Blackwell, D. S. Hibbett, T. Y. James, J. W. Spatafora, and J. W.
Taylor. 2009. The fungi. Current Biology 19:R840–R845.
Stone, M., 2010. Virulent new strains of rust fungus endanger world wheat. Microbe 5(10):423–428.
Strange, R. N., and P. R. Scott. 2005. Plant disease: A threat to global food security. Annual Review
of Phytopathology 43:83–116.
Stuart, S. N., J. S. Chanson, N. A. Cox, B. E. Young, A. Rodrigues, D. L. Fischman, and R. W.
Waller. 2004. Status and trends of amphibian declines and extinctions worldwide. Science
306:1783–1786.
Stuckenbrock, E. H., and B. A. McDonald. 2008. The origins of plant pathogens in agro-ecosystems.
Annual Review of Phytopathology 46:75–100.
Suh, S. O., and M. Blackwell. 2006. Three new asexual arthroconidial yeasts, Geotrichum carabi-
darum sp. nov., Geotrichum histeridarum sp. nov., and Geotrichum cucujoidarum sp. nov.,
isolated from the gut of insects. Mycological Research 110:220–228.
Suh, S. O., C. J. Marshall, J. V. McHugh, and M. Blackwell. 2003. Wood ingestion by passalid beetles
in the presence of xylose-fermenting gut yeasts. Molecular Ecology 12:3137–3145.
Suh, S. O., J. V. McHugh, D. D. Pollock, and M. Blackwell. 2005. The beetle gut: A hyperdiverse
source of novel yeasts. Mycology Research 3:261–265.
Tatum, L. A. 1971. The Southern corn leaf blight epidemic. Science 171:1113–1116.
Taylor, J. W., J. Spatafora, K. O’Donnell, F. Lutzoni, T. James, D. S. Hibbett, D. Geiser, T. D. Bruns,
M. Blackwell. 2004. The Fungi. In Assembling the Tree of Life, edited by J. Cracraft and M. J.
Donoghue. Oxford University Press. Pp. 171–194.
Trethowan, R. M., D. Hodson, H.-J. Braun, W. H. Pfeiffer, and M. Van Ginkel. 2005. Wheat breeding
environments. In Impacts of international wheat breeding research in the developing world,
1988–2002, edited by M. A. Lantican, H. J. Dubin, and M. L. Morris. Mexico. D.F.: CIMMYT.
P. 5.
Tsang, C. A., S. M. Anderson, S. B. Imholte, L. M. Erhardt, S. Chen, B. J. Park, C. Christ, K. K.
Komatsu, T. Chiller, and R. H. Sunenshine. 2010. Enhanced surveillance of coccidioidomycosis,
Arizona, USA, 2007–2008. Emerging Infectious Diseases 16(11):1738–1744.
98 FUNGAL DISEASES
Tyler, B. M., S. Tripathy, X. Zhang, P. Dehal, R.H. Jiang, A. Aerts, F. D. Arredondo, L. Baxter,
D. Bensasson, J. L. Beynon, J. Chapman, C. M. Damasceno, A. E. Dorrance, D. Dou, A. W.
Dickerman, I. L. Dubchak, M. Garbelotto, M. Gijzen, S. G. Gordon, F. Govers, N.J. Grunwald,
W. Huang, K. L Ivors, R.W. Jones, S. Kamoun, K. Krampis, K. H. Lamour, M. K. Lee, W. H.
McDonald, M. Medina, H. J. Meijer, E.K. Nordberg, D. J. Maclean, M. D. Ospina-Giraldo,
P. F. Morris, V. Phuntumart, N. H. Putnam, S. Rash, J. K. Rose, Y. Sakihama, A. A. Salamov,
A. Savidor, C. F. Scheuring, B. M. Smith, B. W. Sobral, A. Terry, T. A. Torto-Alalibo, J.
Win, Z. Xu, H. Zhang, I. V. Grigoriev, D. S. Rokhsar, and J. L. Boore. 2006, Phytophthora
genome sequences uncover evolutionary origins and mechanisms of pathogenesis. Science
313(5791):1261–1266.
Ullstrup, A. J. 1972. The impacts of the Southern corn leaf blight epidemics of 1970–1971. Annual
Reviews in Phytopathology 10:37–50.
Václavík, T. A., E. M. Kanaskie, J. L. Hansen, J. L. Ohmann, and R. K. Meentemeyer. 2010. Pre-
dicting potential and actual distribution of sudden oak death in Oregon: Prioritizing landscape
contexts for early detection and eradication of disease outbreaks. Forest Ecology and Manage-
ment 260:1026–1035.
Vinh, D. C., F. Masannat, R. B. Dzioba, J. N. Galgiani, and S. M. Holland. 2009. Refractory dis-
seminated coccidioidomycosis and mycobacteriosis in inteferon-gamma receptor 1 deficiency.
Clinical Infectious Diseases 49:e62–e65.
Voelz, K. and R. C. May. 2010. Cryptococcal interactions with the host immune system. Eukaryotic
Cell 9:835–846.
Vredenburg, V. T., R. A. Knapp, T. S. Tunstall, and C. J. Briggs. 2010. Dynamics of an emerging dis-
ease drive large-scale amphibian population extinctions. Proceedings of the National Academy
of Sciences, USA (published ahead of print May 10, 2010).
Vurro, M., B. Bonciani, and G. Vannacci. 2010. Emerging infectious diseases of crop plants in
developing countries: Impact on agriculture and socio-economic consequences. Food Security
2(2):113–132.
Wade, N. 1999. For leaf-cutter ants, farm life isn’t so simple. The New York Times, August 3.
Wainwright, M. 1992. The impact of fungi on environmental biogeochemistry. In The Fungal Com-
munity: Its Organization and Role in the Ecosystem, edited by G. C. Carroll and D. T. Wicklow.
New York: Marcel Decker, Inc. Pp. 601-616.
Wake, D. B., and V. T. Vredenburg. 2008. Are we in the midst of the sixth mass extinction? A
view from the world of amphibians. Proceedings of the National Academy of Sciences, USA
105(Suppl 1):11466–11473.
Walker, S. F., J. Bosch, T. Y. James, A. P. Litvintseva, J. A. Oliver Valls, S. Piña, G. García, G. A.
Rosa, A. A. Cunningham, S. Hole, R. Griffiths, and M. C. Fisher. 2008. Invasive pathogens
threaten species recovery programs. Current Biology 18(18):R853–854.
Walker, S. F., J. Bosch, V. Gomez, T. Garner, A. A. Cunningham, D. S. Schmeller, M. Ninyerola, D.€A.
Henk, C. G. Christian-Phillipe Arthur, and M. C. Fisher. 2010. Factors driving pathogenicity vs.
prevalence of amphibian panzootic chytridiomycosis in Iberia. Ecology Letters 2–11.
Warnock, D. W. 2006. Fungal diseases: An evolving public health challenge. Medical Mycology
44(8):697–705.
Weidensaul, S. 2000. The rarest of the rare. Smithsonian 31(8):118–127.
Weinberg, J. 2005. Surveillance and control of infectious diseases at local, national and international
levels. Clinical Microbiology and Infection 11:12–14.
Weir, A., and G. W. Beakes. 1995. An introduction to the Laboulbeniales: A fascinating group of
entomogenous fungi. Mycologist 9:6–10.
Weldon, C., L. H. Du Preez, A. D. Hyatt, R. Muller, and R. Speare. 2004. Origin of the amphibian
chytrid fungus. Emerging Infectious Diseases 10(12):2100–2105.
Weldon, C., A. De Villiers, and L. H. Du Preez. 2007. Quantification of the trade in Xenopus laevis
from South Africa, with implications for biodiversity conservation. African Journal of Herpe-
tology 56(1):77–83.
WORKSHOP OVERVIEW 99
Weldon, C., L. D. Preez, and M. Vences. 2008. Lack of detection of the amphibian chytrid fungus
(Batrachochytrium dendrobatidis) in Madagascar. Monografie del Museo Regionale di Scienze
Naturali di Torino, XLV (2008):95–106.
Wellings, C. R. 2007. Puccinia striiformis in Australia: A review of the incursion, evolution, and
adaptation of stripe rust in the period 1979–2006. Australian Journal of Agricultural Research
58(6):567–575.
WHO (World Health Organization). 2008. International health regulations (2005), 2nd ed. http://
whqlibdoc.who.int/publications/2008/9789241580410_eng.pdf (accessed, May 3, 2011).
_____. 2010. Emerging diseases. http://www.who.int/topics/emerging_diseases/en. (accessed March
9, 2011).
Wibbelt, G., A. Kurth, D. Hellmann, M. Weishaar, A. Barlow, M. Veith, J. Pruger, T. Gorfol, L.
Grosche, F. Bontadina, U. Zophel, H. P. Seidl, P. M. Cryan, and D. S. Blehert. 2010. White-
nose syndrome fungus (Geomyces destructans) in bats, Europe. Emerging Infectious Diseases
16(8):1237–1243.
Wilson, M. E. 2003. The traveler and emerging infections: Sentinel, courier, transmitter. Journal of
Applied Microbiology 94(Suppl):1S–11S.
Woodham-Smith, C. 1962. The great hunger. New York: Harper & Row.
Woodhams, D. C, R. A. Alford, and G. Marantelli. 2003. Emerging disease of amphibians cured by
elevated body temperature. Diseases of Aquatic Organisms 55:65–67.
Woolhouse, M., and E. Gaunt. 2007. Ecological origins of novel human pathogens. Critical Reviews
in Microbiology 33(4):231–242.
Woolhouse, M. E. J., and S. Gowtage-Sequeria. 2005. Host range and emerging and reemerging
pathogens. Emerging Infectious Diseases 11(12):1842–1847.
Xue, C., Y. Tada, X. Dong, and J. Heitman. 2007. The human fungal pathogen cryptococcus can com-
plete its sexual cycle during a pathogenic association with plants. Cell and Microbe 1:263–273.
Zhdanova, N. N., V. A. Zakharchenkoa, V. V. Vembera, and L. T. Nakonechnaya. 2000. Fungi from
Chernobyl: Mycobiota of the inner regions of the containment structures of the damaged nuclear
reactor. Mycological Research 104 (12):1421–1426.
Appendix A
Contributed Manuscripts
A1
1╛╛Reprinted
with kind permission from Springer Science+Business Media: Current Infectious Dis-
eases Reports, The emergence of Cryptococcus gattii in British Columbia and the Pacific Northwest,
10, 2008, p. 108–115, Karen H. Bartlett, Sarah E. Kidd, and James W. Kronstad.
╛╛CurrentInfectious Disease Reports 2008, 10:58–65
╛╛CurrentMedicine Group LLC ISSN 1523-3847
╛╛Copyright © 2008 by Current Medicine Group LLC
James W. Kronstad, PhD, The Michael Smith Laboratories, University of British Columbia, 2185 East
Mall, Vancouver, BC, V6T 1Z4, Canada. Email: kronstad@interchange.ubc.ca.
101
102 FUNGAL DISEASES
Introduction
The basidiomycetous yeast Cryptococcus neoformans has a global distribu-
tion and has achieved prominence in recent decades because of its propensity
to infect immunocompromised people (Casadevall and Perfect, 1998). In fact,
cryptococcosis is recognized as an AIDS-defining illness, and in the absence of
highly active antiretroviral therapy, the disease is a significant cause of death in
individuals with HIV infection (Bicanic and Harrison, 2005; Bicanic et al., 2005).
People and animals acquire the fungus via the inhalation of desiccated yeast
cells or basidiospores from environmental sources such as avian guano, soil, and
trees. Pulmonary infection often results in dissemination to the central nervous
system and C. neoformans is the leading cause of fungal meningitis (Casadevall
and Perfect, 1998).
Isolates of C. neoformans have previously been divided into three varieties
known as grubii, neoformans, and gattii and into serotypes (A–D and hybrids
such as AD) defined by antigenic differences in the capsular polysaccharide that
is the major virulence factor (Casadevall and Perfect, 1998). The gattii variety is
now recognized as a separate species based on phenotypic and molecular traits,
and mating (Kwon-Chung et al., 2002). Thus the current view is that the species
C. neoformans (var grubii and neoformans) contains strains of serotypes A, D,
and AD, and the distinct species C. gattii contains isolates of the B and C sero-
types (Kwon-Chung and Varma, 2006). An excellent review of the differences be-
tween C. gattii and C. neoformans has been published by Sorrell (Sorrell, 2001).
Extensive surveys have been performed over the past 10 years to characterize
the genotypes and distribution of C. neoformans and C. gattii isolates (Barreto de
Oliveira et al., 2004; Boekhout et al., 2001; Boukhout et al., 1997; Fraser et al.,
2005+; Kidd, 2003; Kidd et al., 2004 ++; Kidd et al., 2005+; Meyer et al., 1999;
Meyer et al., 2003). These surveys used a variety of DNA-based typing methods
to provide detailed classifications of isolates into molecular types. Thus, isolates
of C. neoformans var grubii (serotype A) are represented by the VNI, VNII, and
VNB (Litvintseva et al., 2006) molecular types, var neoformans (serotype D) is
represented by the VNIV type, and isolates of the AD hybrid serotype are the
VNIII type. Four molecular types are recognized for C. gattii isolates (designated
VGI–VGIV) and further divisions within the molecular types have been identified
APPENDIX A 103
FIGURE A1-1╇ Map of the forecasted ecologic niche and region of emergence of C. gattii
Figurepotential,
in British Columbia (BC). The optimal, A1-1.eps and unsuitable ecologic niches of C.
gattii in BC are indicated based on biogeoclimatic
bitmap data for the region (Mak, 2007). Note
that the distribution of human and animal cases and the locations of positive environmental
samples coincide primarily with the optimal ecologic niche. The information on human
and animal cases, and environmental sampling, from Washington (WA) is not included.
(Fraser et al., 2005+; Kidd et al., 2005+; Kidd et al., 2007++). For example, VGII
strains can be further classified into VGIIa and VGIIb subtypes, as well as other
less-well characterized subtypes (Kidd et al., 2004; MacDougall et al., 2007++).
There is currently an intense focus on C. gattii due to the unprecedented
emergence of the VGI, VGIIa, and VGIIb molecular types as primary pathogens
of humans and animals on Vancouver Island in British Columbia (BC) (Kidd
et al., 2004; MacDougall et al., 2007++) (Fig. A1-1). Remarkably, the major-
ity of human cases have occurred in people without recognized immunologic
defects, thus highlighting the unusual pathogenicity of C. gattii relative to C.
neoformans. The purpose of this review is to summarize recent progress in the
investigation of this fascinating emergence with regard to human and animal
exposure, environmental colonization, isolate characterization, and the potential
for further dispersal.
104 FUNGAL DISEASES
soil, water, vehicles, and shoes can act as dispersal mechanisms for the organ-
ism (Kidd et al., 2007a++). These mechanisms are consistent with the findings
of a veterinary case-control study, where statistically significant risk factors for
disease in cats and dogs related to soil disturbance within 10 km of cases, log-
ging within 10 km, travel to Vancouver Island, or owner hiking within 6 months
of diagnosis (Duncan et al., 2006c). Although limited environmental sampling
in the San Juan Islands, Olympic Peninsula, and Oregon has not yielded C. gat-
tii (Fraser et al., 2006; Kidd et al., 2007b++; Upton et al., 2007+). Kidd et al.
(2007a++,2007b++) reported finding positive environmental samples from is-
lands in the Georgia Strait and in northern Washington.
A rather surprising finding was that co-isolated C. gattii strains are heteroge-
neous. The first isolates distributed to the research community were mostly from
one sampling site (central Vancouver Island) and may have unduly influenced
our thinking about the composition of the BC outbreak strains (Kidd et al.,
2004++; Fraser et al., 2005+; Fraser et al., 2003). In the initial analysis of the C.
gattii isolates from this site, Kidd et al. (Kidd et al., 2004++) used polymerase
chain reaction (PCR)-fingerprinting to demonstrate that 5% represented the VGI
molecular type and 95% belonged to VGII (90% of these were VGIIa and 10%
were VGIIb based upon a one polymorphic band in the PCR-fingerprint profiles).
Subsequent work revealed that the composition of the C. gattii population var-
ies in different regions where detailed molecular subtyping of isolates has been
undertaken. In the southern extreme of Vancouver Island, VGIIa accounts for
91% of the isolates and the remainder are VGIIb, whereas at another site VGIIa
accounts for only 66% of the isolates, with VGIIb and VGI at 19% and 15%,
respectively (Bartlett and Kidd, unpublished data). Of course, the genotype
frequencies are likely to be dynamic, and repeated sampling is important. Also,
additional diagnostic tools sensitive enough to detect and differentiate isolates
directly in environmental samples (eg, PCR on soil samples) would facilitate a
better understanding of the population structure and mechanisms of spread of the
organism. Already heightened awareness of changing ecologic niches has resulted
in an expansion of knowledge of the environmental origins of other cryptococcal
species (Filion et al., 2006).
interest from these studies was the identification of isolates from other areas of
the world with identical or similar genotypes to the VGIIa (as represented by
isolate A1MR265) and VGIIb (represented by isolate A1MR272) strains from
Vancouver Island. For example, the VGIIa genotype was also shared by the
NIH444 strain (from a patient in Seattle, ca 1971), CBS7750 (from a Eucalyp-
tus tree in San Francisco, ca 1990) and with isolates from other parts of North
America (KB10455 and KB9944) (Fraser et al., 2005+; Kidd et al., 2005+). A
Brazilian clinical isolate, ICB107, differed from the VGIIa genotype at only one
of 22 loci (Fraser et al., 2005+). The VGIIb genotype was also observed among
environmental isolates from Australia (eg, Ram002, Ram005, WM1008), clini-
cal isolates from Australia (eg, NT-6, NT-13), as well as a clinical isolate from
Thailand (MC-S-115) (Fraser et al., 2005+; Kidd et al., 2005+). A Caribbean
strain 99/473 of the VGIIb type was also found to differ at only one of 22 loci
(Fraser et al., 2005+). Intriguingly, two isolates from human cases in Oregon
(2004) were recently found to represent subtypes within the VGII genotype that
have not identified among any other strains to date (MacDougall et al., 2007++).
The VGIIa and VGIIb isolates from Vancouver Island have been obtained
from both clinical and environmental sources. However, the situation is more
complex for strains of the VGI genotype from clinical and environmental sources.
Specifically, Kidd et al. (2005+) characterized six VGI isolates from Vancouver
Island and identified four different genotypes by MLST analysis. Two of these
were environmental isolates with a different genotype from the clinical isolates.
Thus, in contrast to the VGII types, it was not possible to establish an epidemio-
logic link between environmental and clinical isolates of the VGI type. However,
recent analysis of further environmental VGI isolates from Vancouver Island
indicated that they were highly similar to a porpoise isolate (A1MF2863), being
identical at four MLST loci (Kidd and Bartlett, unpublished data). It is possible
that the clinical isolates of the VGI type represent strains acquired during travel
outside of Vancouver Island.
Overall, Kidd et al. (2005+) found that the Vancouver Island isolates were
part of a predominately clonal population with little evidence of sexual recom-
bination occurring between them. Fraser et al. (2005+) also presented evidence
that the VGIIa and VGIIb strains from Vancouver Island were related in that they
shared 14 identical loci out of the 30 examined and proposed that the genotypes
represent either siblings arising from a past mating event, or that one may be the
parent of the other, perhaps as the result of same-sex mating between MATα par-
ents. Selected isolates from Vancouver Island and other parts of the world have
been tested for mating competence. These studies revealed that the VGII isolates
are generally fertile whereas VGI strains are not (Campbell et al., 2005; Fraser
et al., 2003; Kidd et al., 2004++). In general, the ability of C. gattii isolates to
mate has implications for recombination events that might generate strains with
different virulence properties and environmental adaptability.
108 FUNGAL DISEASES
et al., 2007a++), especially with regard to the need for extensive multisource
sampling over many years. The wide distribution of C. gattii genotypes should
also be considered in light of recent reports that infections with this species are
occurring in patients with AIDS (South Africa [Morgan et al., 2006], Southern
California [Chaturvedi et al., 2005a]). Therefore, it will be important to identify
the endemic areas for specific C. gattii genotypes in order to monitor human and
animal disease.
Clinical considerations
Perhaps the most relevant topics regarding the emergence of C. gattii have to
do with identifying risk factors for people, designing ways to limit exposure, and
developing effective methods to treat the infections that do occur. It is common
to see statements in the literature that C. gattii is a primary pathogen that infects
immunocompetent people, and that C. neoformans is an opportunistic pathogen
that infects immunocompromised people. The distinction may be less clear given
that C. gattii is now being found in AIDS patients and C. neoformans can infect
seemingly immunocompetent people (Chaturvedi et al., 2005a; Morgan et al.,
2006; Speed and Dunt, 1995). There is clearly a need for retrospective studies
of patients to determine host risk factors as well as prospective case studies to
determine efficacy of treatments. The number of cases continuing to occur on
Vancouver Island (and among tourists [Lindberg et al., 2007]) would allow this
type of investigation.
An interesting consideration in terms of exploring possible virulence differ-
ences for C. gattii versus C. neoformans is whether mouse virulence studies have
relevance for human disease. For example, the strains with the VGIIa and VGIIb
genotypes from Vancouver Island both cause disease in humans, but laboratory
studies revealed virulence differences between the two strains tested (Fraser et al.,
2005+). The more virulent strain, A1MR265, of the VGIIa genotype showed
equal virulence in the mouse model to strain H99 that is representative of the
most common VNI type of C. neoformans (var grubii). It is possible that these re-
sults reflect the fact that only one isolate of each genotype from Vancouver Island
APPENDIX A 111
was tested and the isolates selected may not be representative. It is clear, however,
that strains of C. gattii show virulence differences (Kronstad, unpublished data)
(Chaturvedi et al., 2005b; Fraser et al., 2005+) and that multiple isolates from
Vancouver Island and worldwide collections need to be tested. The same is true
for C. neoformans as demonstrated by the range of virulence detected by Clancy
et al. (2006). Thus, we need to develop better models to assess differences in
virulence and to explore possible differences that may be relevant to infection of
immunocompetent versus immunocompromised hosts.
ment through the Eucalypts (United Press International, 2007), even though no
link to Eucalyptus was shown in the BC experience (Kidd et al., 2007a++). In
an examination of press coverage of C. gattii as an emerging infectious disease
agent, researchers at the University of BC Centre for Health and Environment
Research found that during the period 2001 to 2006, BC newspapers carried 422
articles warning the public about West Nile Virus (although no West Nile Virus
cases have been reported in BC) compared with 79 articles about C. gattii (170
human cases, eight deaths) (Nicol et al., unpublished data). The research group
concluded that because West Nile Virus is a public health risk with identifiable
precautionary actions in central Canada, newspapers were more likely to print
stock West Nile Virus stories. C. gattii was seen to be a local phenomenon with no
identifiable risk aversion strategies and to have potential economic repercussions
to the areas affected and so was less reported. There also seemed to be confusion
by news writers about the biology of Cryptococcus because the term “virus”
seems to be better understood as a pathogen compared to “yeast” (Nicol et al., un-
published data). Similarly, some news items labeled C. gattii as an “Australian”
fungus despite the body of literature cited above on the global distribution of the
pathogen. Overall, these observations demonstrate that effective education of the
media and the public is a critical component of the management of an emerging
infectious disease.
Conclusions
A great deal has been learned about the emergence of C. gattii in BC over
the past 8 years. We now have a clear picture of the environmental sources of
the pathogen and mechanisms of dispersal, we have an understanding of the
genotypes that are causing disease in humans and animals, and we have some
information about clinical presentation and treatment. Certainly, there is a great
deal more to investigate in terms of risk factors for the human population and
treatment outcomes. In this regard, the situation on Vancouver Island presents an
opportunity to develop a detailed view of an emerging infectious disease with
regard to environmental exposure, the role of sentinel animals in monitoring risk,
and the underlying factors that influence human susceptibility. This information
may prove useful for other emerging diseases and provide methods to manage
both the ongoing situation in BC and the apparent emergence of the disease in
the Pacific Northwest.
Acknowledgments
The authors thank the members of the BC Cryptococcal Working Group
(http://www.cher.ubc.ca/cryptococcus/) and the BC Centre for Disease Control
(http://www. bccdc.org/) for helpful discussions and Sunny Mak for the prepa-
ration of Figure A1-1. The authors are supported in part by grants from the US
APPENDIX A 113
National Institute of Allergy and Infectious Disease (Dr. Kronstad, award RO1-
AI-053721), the Canadian Institutes of Health Research (Drs. Kronstad and
Bartlett), British Columbia Lung Association (Dr. Bartlett), and WorkSafe BC
(Dr. Bartlett). Dr. Kronstad is a Burroughs Wellcome Fund Scholar in Molecular
Pathogenic Mycology, and Dr. Bartlett is a Michael Smith Foundation for Health
Research Scholar.
+ Fraser JA, Giles SS, Wenink EC, et al.: Same-sex mating and the origin of the Vancouver Island
Cryptococcus gattii outbreak. Nature 2005, 437:1360–1364.
An extensive MLST analysis of C. gattii isolates from Vancouver Island and from around the
world. The authors found shared genotypes between the VGIIa and VGIIb strains from BC and
strains of these molecular types from other parts of the world. This study presents interesting
hypotheses about the origin of the VGIIa genotype in BC and reports the first virulence tests of
VGIIa and VGIIb strains from Vancouver Island.
Fraser JA, Lim SM, Diezmann S, et al.: Yeast diversity sampling on the San Juan Islands reveals
no evidence for the spread of the Vancouver Island Cryptococcus gattii outbreak to this locale.
FEMS Yeast Res 2006, 6:620–624.
Fraser JA, Subaran RL, Nichols CB, Heitman J: Recapitulation of the sexual cycle of the primary
fungal pathogen Cryptococcus neoformans var. gattii: implications for an outbreak on Vancou-
ver Island, Canada. Eukaryot Cell 2003, 2:1036–1045.
Hoang LM, Maguire JA, Doyle P, et al.: Cryptococcus neoformans infections at Vancouver Hospital
and Health Sciences Centre (1997–2002): epidemiology, microbiology and histopathology. J
Med Microbiol 2004, 53:935–940.
Kidd SE: Molecular epidemiology and characterization of genetic structure to assess speciation
within the Cryptococcus neoformans complex [PhD thesis]. Sydney: University of Sydney;
2003.
++ Kidd SE, Chow Y, Mak S, et al.: Characterization of environmental sources of the human and
animal pathogen Cryptococcus gattii in British Columbia, Canada, and the Pacific Northwest
of the United States. Appl Environ Microbiol 2007a, 73:1433–1443.
This important study describes a systematic and thorough investigation of the environmental
colonization of C. gattii on Vancouver Island and the Pacific Northwest. Key findings include
the isolation of the pathogen from air, trees, soil, freshwater, and seawater, and the identification
of colonization hotspots. Additionally, this study identified characteristics of soil that may favor
C. gattii colonization.
++ Kidd SE, Bach PJ, Hingston AO, et al.: Cryptococcus gattii dispersal mechanisms, British Colum-
bia, Canada. Emerg Infect Dis 2007b, 13:51–57.
This study employed systematic environmental sampling strategies to document patterns of C.
gattii colonization on Vancouver Island and to obtain evidence for human-mediated dispersal
of the fungus.
+ Kidd SE, Guo H, Bartlett KH, et al.: Comparative gene genealogies indicate that two clonal lineages
of Cryptococcus gattii in British Columbia resemble strains from other geographical areas.
Eukaryot Cell 2005, 4:1629–1638.
This study employed MLST analysis and gene genealogy to reveal a predominantly clonal popu-
lation among the Vancouver Island isolates and to demonstrate that the genotypes of isolates
from BC resembled those of strains from other parts of the world.
++ Kidd SE, Hagen F, Tscharke RL, et al.: A rare genotype of Cryptococcus gattii caused the cryp-
tococcosis outbreak on Vancouver Island (British Columbia, Canada). Proc Natl Acad Sci USA
2004, 101:17258–17263.
This paper describes the results of the first marshaling of the expertise of the international
research community to tackle the analysis of the emergence of C. gattii in BC. The investiga-
tors described initial studies on the environmental source of the pathogen and identified the
molecular types of C. gattii that were responsible for the human and animal cases.
Krockenberger MB, Canfield PJ, Malik R: Cryptococcus neoformans in the koala (Phascolarctos
cinereus): colonization by C n var gattii and investigation of environmental sources. Med Mycol
2002, 40:263–272.
Kwon-Chung KJ, Bennett JE: Epidemiologic differences between the two varieties of Cryptococcus
neoformans. Am J Epidemiol 1984, 120:123–130.
APPENDIX A 115
Kwon-Chung KJ, Boekhout T, Fell JW, Diaz M: (1557) Proposal to conserve the name Cryptococcus
gattii against C. hondurianus and C. bacillisporus (Basidiomycota, Hymenomycetes, Tremel-
lomycetidae). Taxon 2002, 51:804–806.
Kwon-Chung KJ, Varma A: Do major species concepts support one, two or more species within
Cryptococcus neoformans? FEMS Yeast Res 2006, 6:574–587.
Lester SJ, Kowalewich NJ, Bartlett KH, et al.: Clinicopathologic features of an unusual outbreak of
cryptococcosis in dogs, cats, ferrets, and a bird: 38 cases (January to July 2003). J Am Vet Med
Assoc 2004, 225:1716–1722.
Lindberg J, Hagen F, Laursen A, et al.: Cryptococcus gattii risk for tourists visiting Vancouver Island,
Canada. Emerg Infect Dis 2007, 13:178–179.
Litvintseva AP, Thakur R, Vilgalys R, Mitchell TG: Multilocus sequence typing reveals three genetic
subpopulations of Cryptococcus neoformans var grubii (serotype A) including a unique popula-
tion in Botswana. Genetics 2006, 172:2223–2238.
MacDougall L, Fyfe M: Emergence of Cryptococcus gattii in a novel environment provides clues to
its incubation period. J Clin Microbiol 2006, 44:1851–1852.
++ MacDougall L, Kidd SE, Galanis E, et al.: Spread of Cryptococcus gattii in British Columbia,
Canada, and detection in the Pacific Northwest, USA. Emerg Infect Dis 2007, 13:42–50.
This paper describes the detection of C. gattii in three people and eight animals without a travel
history to Vancouver Island, and the detection of the pathogen in air, soil, water and on trees
from sites off the island. The study also reported locally acquired C. gattii infections in three
cats in Washington and two people in Oregon; interestingly, the genotypes of the strains from
the Oregon cases were VGIIa- and VGIIb-like, but MLST results indicated differences from the
isolates of the corresponding subtypes from Vancouver Island.
Mak S: Ecological niche modeling of Cryptococcus gattii in British Columbia [MSc thesis]. Vancou-
ver: University of British Columbia; 2007.
Meyer W, Castaneda A, Jackson S, et al.: Molecular typing of IberoAmerican Cryptococcus neofor-
mans isolates. Emerg Infect Dis 2003, 9:189–195.
Meyer W, Kaocharoen S, Trills L, et al.: Global molecular epidemiology of Cryptococcus gattii VGII
isolates traces the origin of the Vancouver Island outbreak to Latin America [abstract]. Presented
at the 24th Fungal Genetics Conference. Pacific Grove, CA; March 20–25, 2007.
Meyer W, Marszewska K, Amirmostofian M, et al.: Molecular typing of global isolates of Cryp-
tococcus neoformans var neoformans by PCR-fingerprinting and RAPD—a pilot study to
standardize techniques on which to base a detailed epidemiological survey. Electrophoresis
1999, 20:1790–1799.
++ Michael Smith Genome Sciences Center: Cryptococcus Neoformans Summary. http://www.bcgsc.
ca/project/cryptococcus/summary/. Accessed July 9, 2007.
The sequences of the genomes of VGI and VGIIa strains are exceptional resources for detailed
investigations of the virulence properties of C. gattii. In addition, the sequences allow genome-
wide comparative studies with the genomes of C. neoformans var neoformans strains and a var
grubii strain.
Morgan J, McCarthy KM, Gould S, et al.: Cryptococcus gattii infection: characteristics and epidemi-
ology of cases identified in a South African province with high HIV seroprevalence, 2002–2004.
Clin Infect Dis 2006, 43:1077–1080.
Sorrell TC, Brownlee AG, Ruma P, et al.: Natural environmental sources of Cryptococcus neoformans
var gattii. J Clin Microbiol 1996, 34:1261–1263.
Sorrell TC: Cryptococcus neoformans variety gattii. Med Mycol 2001, 39:155–168.
Speed B, Dunt D: Clinical and host differences between infections with the two varieties of Crypto-
coccus neoformans. Clin Infect Dis 1995, 21:28–34.
Stephen C, Lester S, Black W, et al.: Multispecies outbreak of cryptococcosis on southern Vancouver
Island, British Columbia. Can Vet J 2002, 43:792–794.
++ The Broad Institute: Cryptococcus neoformans Serotype B Database. http://www.broad.mit.edu/
annotation/genome/ cryptococcus_neoformans_b. Accessed July 9, 2007.
116 FUNGAL DISEASES
The sequences of the genomes of VGI and VGIIa strains are exceptional resources for de-
tailed investigations of the virulence properties of C. gattii. The genome sequence of a C.
neoformans var grubii strain is also available at the Broad Institute.United Press International:
GE eucalyptus tree investigation urged. http://www.sciencedaily.com/upi/index.php?feed=Sci
ence&article=UPI-1-20070614-13565200-bc-us-eucalyptus.xml. Accessed June 17, 2007.
+ Upton A, Fraser JA, Kidd SE, et al.: First contemporary case of human infection with Cryptococcus
gattii in Puget Sound: evidence for spread of the Vancouver Island outbreak. J Clin Microbiol
2007, In press.
This report and MacDougall et al. (2007) document the recent emergence of C. gattii outside
of BC.
Xu J, Vilgalys R, Mitchell TG: Multiple gene genealogies reveal dispersion and hybridization in the
human pathogenic fungus Cryptococcus neoformans. Mol Ecol 2000, 9:1471–1481.
Zender CS, Talamantes J: Climate controls on Valley Fever incidence in Kern County, California. Int
J Biometeorol 2006, 50:174–182.
A2
Introduction
Human beings were aware of fungal fruiting bodies in prehistoric times, and
the sudden appearance of mushrooms after rain awed those who did not com-
prehend the fungus lifecycle. Lowy (1974) wrote that the sudden appearance of
mushrooms of Amanita muscaria was believed to have been caused by thunder-
3╛╛Louisiana
State University.
4╛╛In
addition to members of Kingdom Fungi, several other organisms of the fungus-like group
Oomycota (Phytophthora) are included.
APPENDIX A 117
bolts as they struck the ground, a belief held independently in Roman, Hindu, and
Mayan cultures. Humans endowed mushrooms with magical properties (Wasson,
1968), and evidence of early fungal use exists in many parts of the world. Grave
guardians, masks, clothing ornaments, and other artifacts were made from the
fruiting bodies of wood-decaying basidiomycetes such as Fomitopsis officinalis
and Haploporus odorus (Blanchette, 1997; Blanchette et al., 1992a, 2002). A
surviving mushroom culture centered on magic mushrooms existed in Oaxaca
for many years, and the celebrated curandera, Maria Sabina, was visited by a
number of prominent individuals and notable musicians who sought her spiritual
guidance (Wasson, 1957, 1976). Although yeasts themselves were not known,
evidence of their activity comes from residues in nine millennia-old Neolithic
vessels (Vouillamoz et al., 2006).
Plant pathogenic fungi also were known in ancient times. Three centuries
BC, Theophrastus recognized fungi as the cause of certain diseases of crops,
but by the first century, the knowledge had been lost, and Pliny attributed lost
yields to the gods or stars (Carefoot and Sprott, 1969). Fungal effects such as
disease were not understood by many until the observations and experiments of
Miles Joseph Berkeley and Anton de Bary around the time of the Irish potato
famine of 1845–1846. This work actually came before the general acceptance
of the germ theory. The contribution of de Bary also argued strongly against
a lingering belief in spontaneous generation (Matta, 2010). Fungi continue to
appear suddenly as they invade natural landscapes to cause diseases of plants
and animals. The invading organisms often are not noticed until they encounter
naïve hosts in new regions where they cause devastating diseases. Earlier inva-
sions included the chestnut blight fungus and several waves of Dutch elm disease
fungi (Alexopoulos et al., 1996). The papers in this volume, Fungal Diseases: An
Emerging Threat to Human, Animal, and Plant Health (IOM, 2011), cover the
newest waves of invasive fungal diseases and their attack on naïve hosts.
More important, however, is the realization that fungi are essential for life on
Earth. Fungi are decomposers that destroy plant and animal bodies and return car-
bon, nitrogen, phosphorus, and other minerals to nutrient cycles. Compatible with
their primary role in decomposition, fungi interact with other living organisms
in nutritional relationships, and their secondary metabolites and enzymes sup-
ply medicines, food and drink, and industrial products for profitable enterprises.
Fungi appear regularly in newspapers and magazines. Over the past year, the
New York Times featured fungi prominently. Articles have included reports of the
identification of a microsporidian fungus partly responsible for colony collapse
disorder of bees, a chytrid responsible for global amphibian decline, Geomyces
destructans of bats, and pathogens of home garden vegetables. Ecological topics
included interactions between bark beetles and fungal symbionts, mycorrhizal as-
sociations, sexual reproduction in truffles, a fungus that exerts selective pressure
on rotifers, and fungal function in the environment. Fungi also have been covered
in the Wall Street Journal as food items, inhabitants of saunas, and the “Torula
118 FUNGAL DISEASES
5╛╛Several brands of skin creams include a variety of basidiomycete fruiting bodies as ingredients
that are said to provide for skin relief and other effects (e.g., Dr. Weil’s Mega-Mushroom lotions,
cleansers, and serums).
6╛╛A group of taxa containing an ancestor and all its descendants.
APPENDIX A 119
the number of DNA markers and taxa in diverse clades to produce increasingly
well-resolved phylogenies,7 the basis of predictive classifications (Figure A2-1)
(Hibbett et al., 2007; James et al., 2006; White et al., 2006). An issue of the jour-
nal Mycologia (98:829–1103, 2006) was devoted to the phylogenetics of many
major groups of fungi. Recent phylogenetic studies have provided new insights
into fungal relationships and show that the earliest diverging lineages of zoo-
sporic8 and zygosporic9 groups are not monophyletic as previously assumed on
the basis of morphological characters and that they are more diverse than previ-
ously understood. Other findings provide data to include microsporidia within or
very near the fungi (Lee et al., 2010). The new phylogenetic studies are largely
the result of several National Science Foundation projects (Research Coordina-
tion Networks: A Phylogeny for Kingdom Fungi [Deep Hypha] and Assembling
the Fungal Tree of Life 1 and 2) that involved more than 100 biologists from
about 20 countries (Blackwell et al., 2006; Hibbett et al., 2007). Current projects
under way include adding taxa to expand the fungal tree of life and pursuing an
increasing number of genomics projects.
About 100,000 species of fungi have been described, but a conservative
estimate suggests that there are 1.5 million fungi on Earth (Hawksworth, 1991,
2001). The estimate has spurred exploration for the million fungi that remain
undiscovered (Figures A2-2A through A2-2D).
More recently the 1.5 million estimate was surpassed by a higher estimate of
3.1 to 5.1 million species based on the use of molecular methods, including high-
throughput sequencing (O’Brien et al., 2005). Because of the great discrepancy
between known and estimated fungal species numbers, mycologists have a re-
newed interest in fungal discovery. Many have wondered, where are the missing
fungi (Hyde, 2001)? If the higher estimates are realistic, the number of fungi is
equal to the number of animal species and may exceed the number of plants by
10:1. Abundant evidence shows that many tropical fungi remain to be discovered
based on species accumulation curves of fungi collected in plots (Aime et al.,
2010). Other habitats reporting large numbers of fungi include living leaves of
tropical trees (Arnold, 2007), soil fungi (O’Brien et al., 2005; Taylor et al., 2010),
and even the fungi in the buildings in which we spend most of our time (Amend
et al., 2010). Many fungi, however, remain to be discovered in northern temperate
regions, including far northeastern Asia (Petersen and Hughes, 2007). We do not
have to look for undescribed fungi in completely new places or tropical regions,
however, because they may be in our backyards. My colleagues and I look for
new species among the yeasts and other microscopic fungi that are difficult to see
a tree diagram.
8╛╛Zoospores are flagellated cells of certain fungi (see Figure A2-1) that are produced in sporangia
in asexual reproduction.
9╛╛Zygospores are thick-walled spores produced in some fungi (see Figure A2-1) resulting from the
“Zygomycetes”
Mucoromycotina ex. Mucor (327 species)
Chytridiomycota
Chytridiomycetales ex. Batrachochytrium (499 spp.)
(706 species)
Spizellomycetales ex. Rhizophlyctis (27 spp.)
FIGURE A2-2╇ Images of representative fungal groups. (A) Hyphae of the blastocladia-
Figure
lean fungus, Allomyces sp. Note A2-2A-D.eps
a terminal zoosporangium (ZS) containing zoospores. The
spiny, dark, thick-walled resting sporesbitmap
(RS) within the hyphae are those of a zoosporic
fungal parasite, Rozella allomycis, of uncertain taxonomic placement. Bar = 10 µm. (B)
Lobosporangium transversale. The zygosporic fungus in the Mortierellales has unusual
spiny lobed sporangia (Benny and Blackwell, 2004). Bar = 50 µm. (C) Sarcoscypha coc-
cinea. The several cm diameter fruiting body of the scarlet cup ascomycete, Sarcoscypha
coccinea. Ascospores are formed within asci on the inner surface of the cup. (D) Mutinus
sp. A stinkhorn similar to one mentioned in the text that caused a search for a dead body
(Anonymous, 2005). Stinkhorns produce noxious compounds that attract insect spore
dispersers. The dark slimy mass of spores has been partially removed by flies.
SOURCES: (A) photo courtesy of Timothy Y. James, provided by Meredith Blackwell
(2009). (B) micrograph courtesy of Kerry O’Donnell, provided by Meredith Blackwell
(2004). (C) photo courtesy of eriotropus/coqui, provided by Meredith Blackwell (2002).
(D) photo courtesy of Nhu H. Nguyen, provided by Meredith Blackwell (2005).
122 FUNGAL DISEASES
with the unaided eye (Boekhout, 2005; Suh et al., 2005), and members of early
diverging lineages that often are difficult to isolate and culture. Ascomycetes and
basidiomycetes are expected to provide the greatest diversity of additional taxa
based on numbers of currently known fungi, but certainly the developing methods
using high-throughput sequencing of DNA will lead to the discovery of more of
the early diverging groups (Figure A2-1) (Kirk et al., 2008).
Examples of large numbers of species isolated into culture from certain sub-
strates include the finding of 418 unique morphotypes of endophytic fungi from
83 leaves in Panama (Arnold, 2007), 257 fungal endophyte genotypes in coffee
plants (Vega et al., 2010), and 650 yeast isolates representing 290 genotypes of
nearly 200 undescribed taxa from the gut of beetles (Suh et al., 2005). Acquiring
cultures and specimens will remain important in cases when fungi and cultures
are needed for certain purposes, including population studies, environmental
remediation, and secondary metabolites. Taylor and his colleagues (2010) used
high-throughput sequencing to estimate the presence of more than 200 taxa in a
0.25 g soil sample with only 14 percent overlap in taxa in a sample taken a meter
away. If we are to determine the number of fungi on Earth, environmental se-
quencing will be necessary to speed fungal exploration and discovery. In addition
to new species, entire lineages, some probably at the level of subphylum, may be
recognized by DNA sequences such as Soil Clone Group 1 (Porter et al., 2008;
Rosling et al., 2010). More work will be needed to determine geographical and
substrate ranges in order to obtain more accurate estimates of species numbers.
Species discovery is relevant to the topic of this workshop because previ-
ously unknown plant and animal pathogenic fungi have been introduced into the
United States many times. These fungi probably caused few symptoms and went
unnoticed in their native hosts. Devastation of naïve hosts, however, led to their
recognition and subsequent description as new species. This scenario certainly
is repeated by the fungi discussed in this meeting, including Batrachochytrium
dendrobatidis, the pathogenic chytrid of amphibians spread around the world;
Geomyces destructans, the pathogen of bats in North America; and Phytophthora
ramorum, causing declines of certain plants in North America and Europe. Prior
invasions have included several fungal agents of Dutch elm disease; the chest-
nut blight fungi; the newly arrived agent of the laurel wilt delivered within the
mycangia10 of its ambrosia beetle vector; and Discula destructans, a pathogen of
North American dogwoods (Alexopoulos et al., 1996; Harrington and Fraedrich,
2010; Zhang and Blackwell, 2001). Recently, a new approach to discovering
the native ranges of certain fungi has been profitable. Ning Zhang (Personal
communication, Rutgers University, December 10, 2010) designed an efficient
assay method using specific primers to detect the dogwood pathogenic fungus in
herbarium specimens. The method promises to greatly reduce the time involved
in determining geographical and host ranges and is ideal for working with col-
10╛╛Mycangia
are pouch-like invaginations in the cuticle of certain insects used to transport cells and
spores of symbiotic fungi, found especially in some species of bark and ambrosia beetles as well as
a few other groups of insects.
APPENDIX A 123
Supermodels
Fungi are important as model systems in research. Saccharomyces cerevisiae
(Figure A2-3) is a supermodel known for its baking and brewing prowess and as
the first eukaryote to have its entire genome sequenced.
FIGURE A2-3╇ Saccharomyces cerevisiae (Y-2235), baker’s yeast and model organism.
Figure
Note the many budding cells in the stained A2-3.eps
preparation. Bar = 5 µm.
bitmapprovided by Meredith Blackwell (2008).
SOURCE: Photo courtesy of Cletus P. Kurtzman,
124 FUNGAL DISEASES
on the border of Italy and Austria. He carried pieces of the fruiting bodies from
two species of wood-rotting basidiomycetes, Piptoporus betulinus and Fomes
fomentarius, perhaps for medicinal uses (Peinter et al., 1998). Other writers
have suggested that one of the fruiting bodies was used as a strop for sharpening
knives and tools, but whatever their use, fungi appear to have been important to
Copper Age Europeans.
Some basidiomycetes have been used medicinally in more recent times.
Extracts of Inonotus obliquus was used in Europe as a treatment for cancer, and
the fruiting bodies of Fomitopsis officinalis (the quinine conk), mentioned earlier
as grave guardians in the Pacific Northwest, were also harvested for medicinal
properties. A different kind of medicinal use by foresters was the application of
sheets of mycelium on ax injuries to stop bleeding (Gilbertson, 1980). The spore
masses of giant puffballs that were discovered stockpiled along Hadrian’s Wall
(in Northern England) also have been used as a styptic (Personal communication,
Roy Watling, former Head of Mycology and Plant Pathology, Royal Botanic Gar-
den Edinburgh, August 27, 1977), and spores of unspecified puffballs also were
widely used as a styptic by natives of North America as well (Blackwell, 2004).
Certain ascomycete fungi, previously known as species of Cordyceps, have
been used in Asian traditional medicine for several centuries (Spatafora et al.,
2007). One of these fungi, a parasite of caterpillars, known as Cordyceps sinensis
since 1878, now is Ophiocordyceps sinensis based on a phylogenetic study (Sung
et al., 2007). Recent interest in the fungus has provided evidence that it may be
effective in the treatment of certain tumors (Spatafora et al., 2007). The revision
of the entire group of insect–pathogenic fungi previously placed in the genus
Cordyceps has resulted in the placement of species in three different families
(Sung et al., 2007). This is an important development because phylogenies are
predictive of traits common to closely related fungi, and other Ophiocordyceps
species may be targeted for the mining of metabolites. The efforts to develop
penicillin for the treatment of bacterial infections at the beginning of World War
II resulted in the discovery of a long-sought magic bullet and hastened the rise
of the modern pharmaceutical industry. In addition to the fungus-derived drug
penicillin, three statin drugs for lowering cholesterol levels (e.g., Lipitor®) and
the immune suppressant cyclosporine each have earned more than a billion dol-
lars annually. Cyclosporine, once critical to transplant surgery, is today used to
treat dry eye as well as more serious conditions (Blackwell, 2011).
Fungi also are big business in the food and beverage industries. In addition to
the usual fresh fruiting bodies of basidiomycetes (mushrooms) and a few highly
favored ascomycetes (truffles and morels), other fungi, such as cuitlacoche (corn
smut) and rice smut, are eaten in Mexico and Asia, respectively. Processed foods
also are made from fungi. These include yeast extract spreads such as marmite
and vegemite and the meat substitute, Quorn™, a product of hyphae of an as-
comycete, a species of Fusarium. Several species of Aspergillus are used in the
processing of soy sauce, and fungi play a part in the flavoring process of cheeses.
126 FUNGAL DISEASES
Throughout the world many fermented foods rely on fungi at least in part to in-
crease nutritional value, improve texture and flavor, and preserve the foodstuff. In
one short street block in Brussels, I examined shop windows to count the many
products that had been touched by fungi: coffee, certain teas, chocolate, cheeses,
bread, salami and dry-cured hams, and numerous fermented beverages (Tamang
and Fleet, 2009). Many African and Asian foods, including miso, ontjom, and
tempeh, are the products of fermentation (Nout, 2009; Rodríguez Couto and
Sanromán, 2006).
As in the case of other fungal products, the making of alcoholic beverages
almost certainly was discovered millennia ago, found accidently in prehistoric
times when wild yeasts settled into a sugary beverage. Yeasts are essential to
the multibillion-dollar alcoholic beverage industry. In the United States, sales
of beer, spirits, and wine were $116 billion in 2003 (Library Index, 2011). The
yeasts involved in brewing were first isolated into pure culture by Emil Hansen
at the Carlsberg Brewery in Copenhagen, and the brewery lab became an im-
portant site of classic yeast genetics and biotechnology research (Hansen and
Kielland-Brandt, 2003). Pretorius (2000) suggested that many additional yeast
species might be used in winemaking. In this context my colleagues and I have
discovered nearly 300 previously unknown yeasts, many of which have the abil-
ity to ferment a variety of sugars, yet are untried for making beverages (Suh et al.,
2005; Urbina and Blackwell, unpublished). In addition to its significance in brew-
ing and bread making, S. cerevisiae, of course, has been extremely important in
industrial biotechnology because of the development of efficient transformation
methods and specialized expression vectors, and for a variety of other genetics
tools (Nevoigt, 2008).
Wood-Decaying Fungi
Fungi are heterotrophic and their ability to degrade organic materials and
return them to nutrient cycles is an essential activity in almost all ecosystems.
The ability of a fungus to degrade specific substrates depends on the enzymes it
produces, and certain fungi are especially important in forest ecosystems where
they are the primary decomposers of wood. Basidiomycetes and some ascomy-
cetes are the primary decomposers of plant cell wall carbohydrates (cellulose and
APPENDIX A 129
collectors and increases the cost of hand-turned bowls at craft fairs. These fungal
effects include the deep blue/green stain of an ascomycete fungus that remains
green in intarsia of fine Italian furniture and the inlay of Tunbridge Ware objects
(Blanchette et al., 1992b). Even Stradivarius violins may have been made more
resonant by the partial decay of the wood (Schwarze et al., 2008).
et al., 2002) and attine ants (Figure A2-6) (Formicidae: Attini) (Mueller et al.,
2005) and of ascomycetes by bark and ambrosia beetles (Scolytinae and Platy-
podinae) (Harrington, 2005). The females of another insect group, siricid wood
wasps (Siricidae), are less well studied, but they have been considered by some
to form farming interactions with fungi (see Gilbertson, 1984). The interaction,
however, does not meet all the criteria established for what has been defined as
“agriculture” (Mueller et al., 2005).
The farming association of the basidiomycete Termitomyces with Old World
macrotermitine termites arose once in Africa. Since that event no additional fun-
gal lineages have been domesticated and no reversals of the fungus to a free-living
state have been found. Repeated host switching, however, has occurred within
termite clades as reflected in the phylogenetic trees of termites and associated
fungi (Aanen et al., 2002). Nest initiation by both males and females of certain
species has been suggested to have influenced the mode of transmission of the
fungus, usually acquired from the environment or some source other than a parent
(horizontal transmission) (Aanen et al., 2002). In the New World it is not termites,
but attine ants that are involved with basidiomycetes in farming interactions, and
Aanen and his colleagues (2002) compared the associations. The attines have
become associated with several clades of fungi, and in contrast to termite trans-
mission, transmission of the fungi is usually directly from parent to offspring
(vertical) except in the early diverging ant lineages. Another important difference
is that the ant-associated fungi apparently do not reproduce sexually. The work
on the fungus–attine ant associations have revealed that ants have evolved with
several groups of fungi on several different occasions. Although the best-known
fungal mutalists are species of Leucocoprinus, other fungal groups, including
certain species of Pterulaceae, have an association with ants in the Apterostigma
pilosum clade (Munkacsi et al., 2004). The intensive studies of the fungi and at-
tine ant associations have led to the discovery of other organisms that participate
in the complex interactions. Species of hypocrealean ascomycetes in the genus
Escovopsis are parasites of the cultivated fungus. Actinomycete associates of the
ants produce antibiotics that have been reported to be specific in inhibiting Es-
covopsis (Currie et al., 1999), but more recently Sen et al. (2009) found that the
bacteria they isolated had more generalized antibiotic activity, including activity
against the cultivated fungus. The association of a fourth component of the as-
sociation is black yeasts that apparently reduce the efficiency of the antibiotics
(Little and Currie, 2008). This attine and—cultivated fungus—Escovopsis para-
site associations provide the best example of coevolution, in this case tripartite
association, among fungi and associates (Currie et al., 2003).
Unlike the termite and ant interactions, fungus-beetle associations have
arisen multiple times. Some bark and ambrosia beetles have mycangia already
mentioned above in which they carry inoculum of certain fungi (Malloch and
Blackwell, 1993). The fungi, often Ceratocystis and Ophiostoma or relatives,
may be the agents of plant diseases, and some of the fungi have been introduced
132 FUNGAL DISEASES
FIGURE A2-6╇Excavation of deeply entrenched nest of the ant Atta texana requires
heavy equipment or, alternatively, ground-penetrating radar to map such nests. The ant is
native to adjacent parts of Texas and Louisiana, and the nests are said to be able to contain
a three-story house. Visual materials based on a ground-penetrating radar nest model are
available on the Internet from Carol LaFayette, Department of Visualization, Texas A&M.
http://www.viz.tamu.edu/faculty/lurleen/main/attatunnel/tunnel.html.
Figure A2-6.eps
SOURCE: Photo courtesy of John Moser, provided by Meredith Blackwell (2009).
bitmap
APPENDIX A 133
with the beetles as in the case of Raffaelea laurelensis, the agent of laurel wilt
disease (Harrington, 2005; Harrington and Fraedrich, 2010). Ophiostoma ulmi
and similar fungi have been introduced into the United States, where they are
virulent pathogens of trees, including American elms. The most efficient dispers-
ers of some of these fungi actually were introduced before the fungus, Ophios-
toma ulmi (Alexopoulos et al., 1996). In this discussion of beneficial fungi, these
interactions benefit the insects and call attention to potential devastating effects
of efficient insect dispersal in the context of emerging plant diseases.
Other beneficial fungal associates of insects involve siricid wood wasps and
wood-decaying basidiomycetes, species of Amylostereum, Stereum, and Daeda-
lea. The wasps lay their eggs through long ovipositors, tube-shaped organs at
the posterior of the abdomen, and the larvae probably rely on fungal enzymes
to decompose and detoxify the wood they ingest (Gilbertson, 1984; Martin,
1992). Many more fungi are associated with insects as necrotrophic parasites
(Figure A2-7), and some of these deadly fungi have potential for development
as biological control agents (Vega et al., 2009). In addition, many of about 1,000
described yeast species have close associations with insects (Table A2-2), and
the yeasts provide important services to the insects (Vega and Dowd, 2005). Cer-
tain clades of gut yeasts appear to have diversified with insect hosts into certain
habitats, and the yeasts provide basic resources for the insects to survive when
subjected to new nutritional situations (Suh et al., 2003, 2006). About 200 species
of Septobasidium in the Septobasidiales are known as associates of scale insects;
only a few related species of Pachnocybe grow on wood (Henk and Vilgalys,
2007). The use of insect hosts is unusual for fungi that are related to the plant
pathogenic rust fungi. The fungi are parasites of a few of the scale individuals,
but in general benefit the entire insect colony by providing a protective covering
against parasitic wasps (Henk and Vilgalys, 2007). Two orders of zygomycetes,
Harpellales and Asellariales, were previously placed in a polyphyletic group
known as Trichomycetes. The results of several studies indicate that these gut
fungi produce vitamins and perhaps other benefits for their aquatic insect hosts
(Lichtwardt et al., 2001). One species is known to parasitize simulid black flies
(Lichtwardt et al., 2001), potentially a benefit to those who engage in outdoor
activities.
Another nutritional interaction between fungi and animals is only briefly
noted here, but is extremely important. An early diverging lineage of obligately
anaerobic multiflagellated fungi, the Neocallimastigomycota, and vertebrate her-
bivores are closely associated (Griffith et al., 2010). The fungi reside in the host
rumen or another anaerobic part of the gut, where they are important in supplying
cellulases and other enzymes for the degradation of the large quantities of cel-
lulose ingested by the herbivore (James et al., 2006).
Conclusion
Many fungi are obligate, beneficial associates of other groups of organisms.
These are the “good fungi” of this article, and we often fail to appreciate their
value because the fungi usually are unseen within their substrates unless they
form macroscopic fruiting bodies. More often it is the effects of the fungi that we
observe when they ferment fruit juice, or fitting to this volume, cause dramatic
new outbreaks of disease. The Robert Frost poem quoted in the prologue of this
publication describes the costs of the introduction of the disease caused by the
chestnut blight fungus, Cryphonectria parasitica. The poem predicts that the
disease will ravage until a new pathogen comes to kill the fungus, and in fact a
virus did appear to suppress the fungus. In 1974, however, yet another pathogen,
the oriental chestnut gall wasp, was introduced to attack the trees, an additional
turn not predicted by the verse.
Today, as one out of every six or seven humans on Earth is reported to be
malnourished or hungry (FAO, 2010), the war against pathogenic diseases of
plants and animals is as important as ever. An earlier writer, Jonathan Swift
(1667–1745) addressed the topic of hunger with his essay, A Modest Proposal,
written to bring attention to the starvation of Irish tenant farmers during the po-
tato famine. In Gulliver’s Travels he wrote directly of the importance of increas-
ing agriculture yields:
APPENDIX A 135
And he gave it for his opinion, “that whoever could make two ears of corn, or
two blades of grass, to grow upon a spot of ground where only one grew before,
would deserve better of mankind, and do more essential service to his country,
than the whole race of politicians put together.”
—Jonathan Swift, Gulliver’s Travels, Part II, Voyage to Brobdingnag,
first published in 1726–1727.
Acknowledgments
I am grateful to Dr. Fernando Vega, who improved the original manuscript
through his careful editing. Several colleagues provided images, and Dr. Matthew
Brown kindly prepared the plate. I acknowledge support from the National Sci-
ence Foundation (NSF-0732671 and DEB-0417180) and the Louisiana State
University Boyd Professor support fund.
References
Aanen, D. K., P. Eggleton, C. Rouland-Lefèvre, T. Guldberg-Frøslev, S. Rosendahl, and J. J.
Boomsma. 2002. The evolution of fungus-growing termites and their mutualistic fungal symbi-
onts. Proceedings of the National Academy of Sciences, USA 99:14887–14892.
Ahmadjian, V. 1993. The lichen symbiosis. New York: John Wiley and Sons.
Aime, M. C., D. L. Largent, T. W. Henkel, and T. J. Baroni. 2010. The entolomataceae of the Pa-
karaima Mountains of Guyana IV: New species of Calliderma, Paraeccilia and Trichopilus.
Mycologia 102:633–649.
Alexopoulos, C. J., C. W. Mims, and M. Blackwell. 1996. Introductory mycology. New York: John
Wiley and Sons.
Amend, A. S., K. A. Seifert, R. Samson, and T. D. Bruns. 2010. Indoor fungal composition is geo-
graphically patterned and more diverse in temperate zones than in the tropics. Proceedings of
the National Academy of Sciences, USA 107:13748–13753.
Anderson, J. B., J. Funt, D. A. Thompson, S. Prabhu, A. Socha, C. Sirjusingh, J. R. Dettman, L.
Parreiras, D. S. Guttman, A. Regev, and L. M. Kohn. 2010. Determinants of divergent adapta-
tion and Dobzhansky-Muller interaction in experimental yeast populations. Current Biology
20:1383–1388.
Anonymous. 2005. “.€.€.€while in Germany, a search for a body turns up mushrooms.” The Mycophile
46 (6):5 http://www.namyco.org/images/pdf_files/MycophileNovDec05.pdf [reprint from Re-
uters Limited web site, Updated: 3:45 p.m. ET Aug. 2, 2005].
Arnold, A. E. 2007. Understanding the diversity of foliar endophytic fungi: Progress, challenges, and
frontiers. Fungal Biology Reviews 21:51–66.
Arnold, A. E., L. C. Mejía, D. Kyllo, E. I. Rojas, Z. Maynard, N. Robbins, and E. A. Herre. 2003.
Fungal endophytes limit pathogen damage in a tropical tree. Proceedings of the National Acad-
emy of Sciences, USA 100:15649–15654.
Beadle, G. W. 1958. Nobel lecture. http://nobelprize.org/nobel_prizes/medicine/laureates/1958/
beadle-lecture.html (accessed February 24, 2011).
136 FUNGAL DISEASES
Benny, G. L., and M. Blackwell. 2004. Lobosporangium, a new name for Echinosporangium Malloch,
and Gamsiella, a new genus for Mortierella multidivaricata. Mycologia 96:143–149.
Bhat, M. K. 2000. Cellulases and related enzymes in biotechnology. Biotechnology Advances 18:
355–383.
Blackwell, M. 2011. The fungi: 1, 2, 3, … 5.1 million species? American Journal of Botany:
98:426–438.
Blackwell, M., D. S. Hibbett, J. W. Taylor, and J. W. Spatafora. 2006. Research coordination networks:
A phylogeny for kingdom Fungi (Deep Hypha). Mycologia 98:829–837.
Blackwell, M., C. P. Kurtzman, M.-A. Lachance, and S.-O. Suh. 2009a. Saccharomycotina. Saccha-
romycetales. Version 22 January 2009. http://tolweb.org/Saccharomycetales/29043/2009.01.22
(accessed March 30, 2011).
Blackwell, M., R. Vilgalys, T. Y. James, and J. W. Taylor. 2009b. Fungi. Eumycota: mush-
rooms, sac fungi, yeast, molds, rusts, smuts, etc. Version 10 April 2009. http://tolweb.org/
Fungi/2377/2009.04.10 (accessed March 30, 2011).
Blackwell, W. H. 2004. Puffballs: Overlooked medicinals? Mushroom, the Journal Fall 2004:1–5.
Blanchette, R. A. 1997. Haploporus odorus: A sacred fungus in traditional Native American culture
of the northern plains. Mycologia 89:233–240.
Blanchette, R. A., B. D. Compton, N. J. Turner, and R. L. Gilbertson. 1992a. Nineteenth century
shaman grave guardians are carved Fomitopsis officinalis sporophores. Mycologia 84:119–124.
Blanchette, R. A., A. M. Wilmering, and M. Baumeister. 1992b. The use of green-stained wood caused
by the fungus Chlorociboria in intarsia masterpieces from the 15th-century. Holzforschung
46:225–232.
Blanchette, R. A., C. C. Renner, B. W. Held, C. Enoch, and S. Angstman. 2002. The current use of
Phellinus igniarius by the Eskimos of Western Alaska. Mycologist 16:142–145.
Boekhout, T. 2005. Gut feeling for yeasts. Nature 434:449–451.
Buchner, P. 1965. Endosymbiosis of animals with plant microorganisms. New York: John Wiley and Sons.
Carefoot, E. R., and G. L. Sprott. 1969. Famine on the wind: Man’s battle against plant disease.
Chicago, IL: Rand McNally and Company.
Currie, C. R., J. A. Scott, R. C. Summerbell, and D. Malloch. 1999. Fungus-growing ants use
antibiotic-producing bacteria to control garden parasites. Nature 398:701–704.
Currie, C. R., B. Wong, A. E. Stuart, T. R. Schultz, S. A. Rehner, U. G. Mueller, G.-H. Sung, J. W.
Spatafora, and N. A. Straus. 2003. Ancient tripartite coevolution in the attine ant-microbe sym-
biosis. Science 299:386–388.
Dowd, P. F. 1991. Symbiont-mediated detoxification in insect herbivores. In Microbial mediation of
plant–herbivore interactions, edited by P. Barbosa, V. A. Krischik, and C. G. Jones. New York:
John Wiley and Sons. Pp. 411–440.
FAO (Food and Agriculture Organization). 2010. The state of food insecurity in the world 2010. http://
www.fao.org/docrep/013/i1683e/i1683e.pdf (accessed June 13, 2011).
Farrell, B. D., A. S. Sequeira, B. C. O’Meara, B. B. Normark, J. H. Chung, and B. H. Jordal. 2001.
The evolution of agriculture in beetles (Curculionidae: Scolytinae and Platypodinae). Evolution
55:2011–2027.
Gilbertson, R. L. 1980. Wood-rotting fungi of North America. Mycologia 72:1–49.
———. 1984. Relationships between insects and wood-rotting basidiomycetes. In Fungus-insect
relationships, perspectives in ecology and evolution, edited by Q. Wheeler and M. Blackwell.
New York: Columbia University Press. Pp. 130–165.
Griffith, G., S. Baker, K. Fliegerova, A. Liggenstoffer, M. van der Giezen, and G. Beakes. 2010.
Anaerobic fungi: Neocallimastigomycota. IMA Fungus 1:181–185.
Hansen, J., and M. C. Kielland-Brandt. 2003. Brewer’s yeast: Genetic structure and targets for
improvement. In Functional genetics of industrial yeasts, edited by J. H. de Winde. Berlin,
Germany: Springer. Pp. 143–170.
Harrington, T. C. 2005. Ecology and evolution of mycophagous bark beetles and their fungal partners.
In Insect–fungal associations: Ecology and evolution, edited by F. E. Vega and M. Blackwell.
New York: Oxford University Press. Pp. 257–291.
APPENDIX A 137
Harrington, T. C., and S. W. Fraedrich. 2010. Quantification of propagules of the laurel wilt fungus
and other mycangial fungi from the redbay ambrosia beetle, Xyleborus glabratus. Phytopathol-
ogy 100:1118–1123.
Hawksworth, D. L. 1991. The fungal dimension of biodiversity: Magnitude, significance, and con-
servation. Mycological Research 95:641–655.
———. 2001. The magnitude of fungal diversity: The 1.5 million species estimate revisited. Myco-
logical Research 105:1422–1432.
Henk, D. A., and R. Vilgalys. 2007. Molecular phylogeny suggests a single origin of insect symbiosis
in the Pucciniomycetes with support for some relationships within the genus Septobasidium.
American Journal of Botany 94:1515–1526.
Hibbett, D. M., M. Binder, J. F. Bischoff, M. Blackwell, P. F. Cannon, O. Eriksson, S. Huhndorf, T. Y.
James, P. M. Kirk, R. Lücking, T. Lumbsch, F. Lutzoni, P. B. Matheny, D. J. McLaughlin, M. J.
Powell, S. Redhead, C. L. Schoch, J. W. Spatafora, J. A. Stalpers, R. Vilgalys, M. C. Aime, A.
Aptroot, R. Bauer, D. Begerow, G. L. Benny, L. A. Castlebury, P. W. Crous, Y.-C. Dai, W. Gams,
D. M. Geiser, G. W. Griffith, D. L. Hawksworth, V. Hofstetter, K. Hosaka, R. A. Humber, K.
Hyde, U. Kõljalg, C. P. Kurtzman, K.-H. Larsson, R. Lichtwardt, J. Longcore, A. Miller, J.-M.
Moncalvo, S. Mozley Standridge, F. Oberwinkler, E. Parmasto, J. D. Rogers, L. Ryvarden, J.
P. Sampaio, A. Schuessler, J. Sugiyama, J. W. Taylor, R. G. Thorn, L. Tibell, W. A. Untereiner,
C. Walker, Z. Wang, A. Weir, M. Weiss, M. White, K. Winka, Y.-J. Yao, and N. Zhang. 2007.
A higher-level phylogenetic classification of the fungi. Mycological Research 111:509–547.
Horsfall, J. G. 1958. The fight with the fungi: The rusts and rots that rob us, the blasts and blights that
beset us. In Fifty years of botany: Golden jubilee volume of the Botanical Society of America,
edited by W. C. Steere. New York: McGraw-Hill. Pp. 50–60.
Horsfall, J. G., and E. B. Cowling. 1978. Some epidemics man has known. In Plant pathology: An
advanced treatise. Vol. 2. The diseased plant, edited by J. G. Horsfall and E. B. Cowling. New
York: Academic Press. Pp. 17–32.
Hyde, K. D. 2001. Where are the missing fungi? Mycological Research 105:1409–1412.
IOM (Institute of Medicine). 2011. Fungal diseases: An emerging challenge to human, animal, and
plant health—a workshop summary. Washington, DC: The National Academies Press.
James, T. Y., P. M. Letcher, J. E. Longcore, S. E. Mozley-Standridge, D. Porter, M. J. Powell, G. W.
Griffith, and R. Vilgalys. 2006. A molecular phylogeny of the flagellated Fungi (Chytridiomy-
cota) and description of a new phylum (Blastocladiomycota). Mycologia 98:860–871.
Jeffries, T. W., I. V Grigoriev, J. Grimwood, J. M. Laplaza, A. Aerts, A. Salamov, J. Schmutz, E.
Lindquist, P. Dehal, H. Shapiro, Y.-S. Jin, V. Passoth, and P. M. Richardson. 2007. Genome
sequence of the lignocellulose-bioconverting and xylose-fermenting yeast Pichia stipitis. Nature
Biotechnology 25:319–326.
Joint Genome Institute. 2007. Super-fermenting fungus genome sequenced. To be harnessed for im-
proved biofuels production. http://www.jgi.doe.gov/News/news_3_5_07.html (accessed March
25, 2011).
Lachance, M. A., W. T. Starmer, C. A. Rosa, J. M. Bowles, J. S. F. Barker, and D. H. Janzen. 2001.
Biogeography of the yeasts of ephemeral flowers and their insects. FEMS Yeast Research 1:1–8.
Lee, S. C., N. Corradi, S. Doan, F. S. Dietrich, P. J. Keeling, and J. Heitman. 2010. Evolution of the
sex-related locus and genomic features shared in Microsporidia and Fungi. PLoS ONE 5:e10539.
Library Index. 2011. U.S. alcohol sales and consumption. http://www.libraryindex.com/pages/2127/
Economics-Alcohol-Tobacco-U-S-ALCOHOL-SALES-CONSUMPTION.html (accessed
March 28, 2011).
Lichtwardt, R. W., M. J. Cafaro, and M. M. White. 2001. The Trichomycetes, fungal associates of
arthropods. Revised edition. http://www.nhm.ku.edu/~fungi/monograph/text/mono.htm (ac-
cessed March 22, 2011).
Little, A. E. F., and C. R. Currie. 2008. Black yeast symbionts comprise the efficiency of antibiotic
defenses in fungus-growing ants. Ecology 89:1216–1222.
Lowy, B. 1974. Amanita muscaria and the thunderbolt legend in Guatemala and Mexico. Mycologia
66:188–191.
138 FUNGAL DISEASES
Malloch, D., and M. Blackwell. 1993. Dispersal biology of ophiostomatoid fungi. In Ceratocystis
and Ophiostoma: Taxonomy, ecology and pathology, edited by M. J. Wingfield, K. A. Seifert,
and J. F. Webber. St. Paul, MN: APS. Pp. 195–206.
Márquez, L. M., R. S. Redman, R. J. Rodriguez, and M. J. Roossinck. 2007. A virus in a fungus in a
plant—three-way symbiosis required for thermal tolerance. Science 315:513–515.
Martin, M. M. 1992. The evolution of insect–fungus associations: From contact to stable symbiosis.
American Zoologist 32:593–605.
Matta, C. 2010. Spontaneous generation and disease causation: Anton de Bary’s experiments with
Phytophthora infestans and late blight of potato. Journal of the History of Biology 43:459–491.
Mewes, H. W., K. Albermann, M. Bähr, D. Frishman, A. Gleissner, J. Hani, K. Heumann, K. Kleine,
A. Maier, S. G. Oliver, F. Pfeiffer, and A. Zollner. 1997. Overview of the yeast genome. Nature
387:7–8.
Mueller, U. G., N. M. Gerardo, T. R. Schultz, D. Aanen, and D. Six. 2005. The evolution of agriculture
in insects. Annual Review of Ecology and Systematics 36:563–569.
Munkacsi, A. B., J. J. Pan, P. Villesen, U. G. Mueller, M. Blackwell, and D. J. McLaughlin. 2004.
Convergent coevolution in the domestication of coral mushrooms by fungus-growing ants.
Proceedings of the Royal Society of London, Series B, Biological Sciences 271:1777–1782.
Nash, T. H. 2008. Lichen biology, 2nd ed. Cambridge, U.K.: Cambridge University Press.
Nevoigt, E. 2008. Progress in metabolic engineering of Saccharomyces cerevisiae. Microbiology and
Molecular Biology Reviews 72:379–412.
Nout, M. J. R. 2009. Rich nutrition from the poorest: Cereal fermentations in Africa and Asia. Food
Microbiology 26:685–692.
O’Brien, B. L., J. L. Parrent, J. A. Jackson, J. M. Moncalvo, and R. Vilgalys. 2005. Fungal commu-
nity analysis by large-scale sequencing of environmental samples. Applied and Environmental
Microbiology 71:5544–5550.
Peinter, U., R. Pöder, and T. Pümpel. 1998. The iceman’s fungi. Mycological Research 102:1153–1162.
Petersen, R. H., and K. W. Hughes. 2007. Some agaric distributions involving Pacific landmasses and
Pacific Rim. Mycoscience 48:1–14.
Porter, T. M., C. W. Schadt, L. Rizvi, A. P. Martin, S. K. Schmidt, L. Scott-Denton, R. Vilgalys,
and J. M. Moncalvo. 2008. Widespread occurrence and phylogenetic placement of a soil clone
group adds a prominent new branch to the fungal tree of life. Molecular Phylogenetics and
Evolution 46:635–644.
Pretorius, I. S. 2000. Tailoring wine yeast for the new millennium: Novel approaches to the ancient
art of winemaking. Yeast 16:675–729.
Rodriguez, R. J., J. F. White, Jr., A. E. Arnold, and R. S. Redman. 2009. Fungal endophytes: Diversity
and functional roles. New Phytologist 182:314–330.
Rodríguez Couto, S., and M. A. Sanromán. 2006. Application of solid-state fermentation to food
industry—a review. Journal of Food Engineering 76:291–302.
Rosling, A., K. Cruz Martinez, A. Menkis, K. Ihrmark, S. Holmström, S. Norström, A. Broberg, and
B. D. Lindahl et al. 2010. Getting to know the fungi in Soil Clone Group 1. Abstract. Interna-
tional Mycological Congress, Edinburgh, Scotland. August 4, 2010.
Schoch, C. L., G.-H. Sung, F. L. López-Giráldez, J. P. Townsend, J. Miadlikowska, V. Rie Hofstetter,
B. Robbertse, P. B. Matheny, F. Kauff, Z. Wang, C. Gueidan, R. M. Andrie, K. Trippe, L. M.
Ciufetti, A. Wynns, E. Fraker, B. P. Hodkinson, G. Bonito, J. Z. Groenewald, M. Arsanlou, G. S.
De Hoog, P. W. Crous, D. Hewitt, D. H. Pfister, K. Peterson, M. Grysenhout, M. J. Wingfield, A.
Aptroot, S.-O. Suh, M. Blackwell, D. M. Hillis, G. W. Griffith, L. A. Castlebury, A. Y. Rossman,
H. T. Lumbsch, R. L. Lücking, B. Büdel, A. Rauhut, P. Diederich, D. Ertz, D. M. Geiser, K.
Hosaka, P. Inderbitzin, J. Kohlmeyer, B. Volkmann-Kohlmeyer, L. Mostert, K. O’Donnell, H.
Sipman, J. D. Rogers, R. A. Shoemaker, J. Sugiyama, R. C. Summerbell, W. Untereiner, P. R.
Johnston, S. Stenroos, A. Zuccaro, P. S. Dyer, P. D. Crittenden, M. S. Cole, K. Hansen, J. M.
Trappe, R. Yahr, F. Lutzoni, and J. W. Spatafora. 2009. The Ascomycota tree of life: A phylum-
wide phylogeny clarifies the origin and evolution of fundamental reproductive and ecological
traits. Systematic Biology 58:224–239.
APPENDIX A 139
Schwarze, F. W., M. Spycher, and S. Fink. 2008. Superior wood for violins—wood decay fungi as a
substitute for cold climate. New Phytologist 179:1095–1104.
Selosse, M.-A., F. Richard, X. He, and S. W. Simard. 2006. Mycorrhizal networks: des liaisons dan-
gereuses? Trends in Ecology and Evolution 21:621–628.
Sen, R., H. D. Ishak, D. Estrada, S. E. Dowd, E. Hong, and U. G. Mueller. 2009. Generalized an-
tifungal activity and 454-screening of Pseudonocardia and Amycolatopsis bacteria in nests of
fungus-growing ants. Proceedings of the National Academy of Sciences, USA 106:17805–17810.
Smith, S. E., and D. J. Read. 2008. Mycorrhizal symbiosis. San Diego, CA: Academic.
Spatafora, J. W., G.-H. Sung, J.-M. Sung, N. Hywel-Jones, and J. F. White. 2007. Phylogenetic
evidence for an animal pathogen origin of ergot and the grass endophytes. Molecular Ecology
16:1701–1711.
Starmer, W. T., R. A. Schmedicke, and M. A. Lachance. 2006. The origin of the cactus-yeast com-
munity. FEMS Yeast Research 3:441–448.
Suh, S.-O., C. J. Marshall, J. V. McHugh, and M. Blackwell. 2003. Wood ingestion by passalid beetles
in the presence of xylose-fermenting gut yeasts. Molecular Ecology 12:3137–3145.
Suh, S.-O., J. V. McHugh, D. Pollock, and M. Blackwell. 2005. The beetle gut: A hyperdiverse source
of novel yeasts. Mycological Research 109:261–265.
Suh, S.-O., M. Blackwell, C. P. Kurtzman, and M.-A. Lachance. 2006. Phylogenetics of Saccharo-
mycetales, the ascomycete yeasts. Mycologia 98:1008–1019.
Sung, G.-H., N. L. Hywel-Jones, J.-M. Sung, J. Luangsa-ard, B. Shrestha, and J. W. Spatafora. 2007.
Phylogenetic classification of Cordyceps and the clavicipitaceous fungi. Studies in Mycology
57:5–59.
Tamang, J. P., and G. H. Fleet. 2009. Yeast diversity in fermented foods and beverages. In Yeast
biotechnology: Diversity and application, edited by T. Satyanarayana and G. Kunze. Berlin,
Germany: Springer. Pp. 169–198.
Tatum, E. L. 1958. Nobel lecture. http://nobelprize.org/nobel_prizes/medicine/laureates/1958/tatum-
lecture.html (accessed June 13, 2011).
Taylor, D. L., I. C. Herriott, K. E. Stone, J. W. McFarland, M. G. Booth, and M. B. Leigh. 2010.
Structure and resilience of fungal communities in Alaskan boreal forest soils. Canadian Journal
of Forest Research 40:1288–1301.
Trappe, J. M. 1987. Phylogenetic and ecologic aspects of mycotrophy in the angiosperms from an
evolutionary standpoint. In Ecophysiology of VA mycorrhizal plants, edited by G. R. Safir. Boca
Raton, FL: CRC Press. Pp. 2–25.
Urbina, H. and M. Blackwell. 2010 (unpublished). Yeasts associated with wood-ingesting beetles.
Baton Rouge, LA: Louisiana State University.
Van Vleet, J. H., and T. W. Jeffries. 2009. Yeast metabolic engineering for hemicellulosic ethanol
production. Current Opinion in Biotechnology 20:300–306.
Vega, F. E., and P. F. Dowd. 2005. The role of yeasts as insect endosymbionts. In Insect–fungal as-
sociations: Ecology and evolution, edited by F. E. Vega and M. Blackwell. New York: Oxford
University Press. Pp. 211–243.
Vega, F. E., M. S. Goettel, M. Blackwell, D. Chandler, M. A. Jackson, S. Keller, M. Koike, N. K.
Maniania, A. Monzón, B. H. Ownley, J. K. Pell, D. E. N. Rangel, and H. E. Roy. 2009. Fungal
entomopathogens: New insights on their ecology. Fungal Ecology 2:149–159.
Vega, F. E., A. Simpkins, M. C. Aime, F. Posada, S. W. Peterson, S. A. Rehner, F. Infante, A. Castillo,
and A. E. Arnold. 2010. Fungal endophyte diversity in coffee plants from Colombia, Hawai’i,
Mexico, and Puerto Rico. Fungal Ecology 3:122–138.
Vouillamoz, J. F., P. E. McGovern, A. Ergul, G. Söylemezoğlu, G. Tevzadze, C. P. Meredith, and
M. S. Grando. 2006. Genetic characterization and relationships of traditional grape cultivars
from Transcaucasia and Anatolia. Plant Genetic Resources 4:144–158.
Wasson, R. G. 1957. Seeking the magic mushroom. Life magazine, May 13, 1957:100–120.
———. 1968. Soma: Divine mushroom of immortality. New York: Harcourt Brace Jovanovich.
———. 1976. Maria Sabina and her Mazatec mushroom velada. New York: Harcourt.
140 FUNGAL DISEASES
Wheeler, Q. D., and M. Blackwell. 1984. Fungus–insect relationships: Perspectives in ecology and
evolution. New York: Columbia University Press.
White, M. M., T. Y. James, K. O’Donnell, M. J. Cafaro, Y. Tanabe, and J. Sugiyama. 2006. Phylogeny
of the Zygomycota based on nuclear ribosomal sequence data. Mycologia 98:872–884.
Wilding, N., N. M. Collins, P. M. Hammond, and J. F. Webber. 1989. Insect–fungus interactions.
New York: Academic Press.
Zhang, N., and M. Blackwell. 2001. Molecular phylogeny of dogwood anthracnose fungus (Discula
destructiva) and the Diaporthales. Mycologia 93:356–364.
A3
Methods
Technological advances make it possible to apply molecular methods to
develop a stable classification and to discover and identify fungal taxa.
11╛╛Reprinted with kind permission from the Botanical Society of America, www.amjbot.org.
12╛╛Manuscript received 10 August 2010; revision accepted 19 January 2011.
╛╛The author thanks N. H. Nguyen, H. Raja, and J. A. Robertson for permission to use their photo-
graphs, two anonymous reviewers who helped to improve the manuscript, and David Hibbett, who
graciously provided an unpublished manuscript. She acknowledges funding from NSF DEB-0417180
and NSF-0639214.
13╛╛Key words: biodiversity; fungal habitats; fungal phylogeny; fungi; molecular methods; numbers
of fungi.
14╛╛Department of Biological Sciences; Louisiana State University; Baton Rouge, Louisiana 70803
USA.
15╛╛Author for correspondence (e-mail: mblackwell@lsu.edu) doi:10.3732/ajb.1000298.
APPENDIX A 141
Key Results
Molecular methods have dramatically increased our knowledge of Fungi in
less than 20 years, revealing a monophyletic kingdom and increased diversity
among early-diverging lineages. Mycologists are making significant advances in
species discovery, but many fungi remain to be discovered.
Conclusions
Fungi are essential to the survival of many groups of organisms with which
they form associations. They also attract attention as predators of invertebrate
animals, pathogens of potatoes and rice and humans and bats, killers of frogs and
crayfish, producers of secondary metabolites to lower cholesterol, and subjects
of prize winning research. Molecular tools in use and under development can be
used to discover the world’s unknown fungi in less than 1000 years predicted at
current new species acquisition rates.
[sic]
Figure A3-1.eps
FIGURE A3-1╇ Fungal phyla and approximate number of species in each group (Kirk
et al., 2008). Evidence from gene order conversion and multilocus sequencing indicates
bitmap
that microsporidians are Fungi (see below; Lee et al., 2010). Note also that zoosporic and
zygosporic fungal groups are not supported as monophyletic. Tree based on Hibbett et al.
(2007), White et al. (2006), and James et al. (2006).
zoosporic and zygosporic lineages than previously realized (Bowman et al., 1992;
Blackwell et al., 2006; Hibbett et al., 2007; Stajich et al., 2009).
Sequences of one or several genes are no longer evidence enough in phylo-
genetic research. A much-cited example of the kind of problem that may occur
when single genes with different rates of change are used in analyses involves
Microsporidia. These organisms were misinterpreted as early-diverging eukary-
otes in the tree of life based on their apparent reduced morphology (Cavalier-
Smith, 1983). Subsequently, phylogenetic analyses using small subunit ribosomal
RNA genes wrongly supported a microsporidian divergence before the origin of
mitochondria in eukaryotic organisms (Vossbrinck et al., 1987). More recent mor-
phological and physiological studies have not upheld this placement, and analy-
ses of additional sequences, including those of protein-coding genes, support the
view that these obligate intracellular parasites of insect and vertebrate hosts are
members of the Fungi (Keeling, 2009; Corradi and Keeling, 2009). Additional
evidence from genome structure as well as phylogenetic analyses, supports the
inclusion of microsporidians within the Fungi and indicates that comparison of
whole genomes contributes to the solution of challenging phylogenetic problems
(Lee et al., 2010).
The level of resolution and sophistication of systematics studies made pos-
sible by molecular markers and phylogenetic analyses put mycologists on equal
footing with other biologists for competitive funding, and they joined in several
community-wide efforts to organize fungal diversity within a phylogenetic classi-
fication. Three projects funded by the National Science Foundation were initiated,
including the Research Coordination Network: A Phylogeny for Kingdom Fungi
(Deep Hypha) and successive Tree of Life projects, Assembling the Fungal Tree
of Life (AFTOL-1) and a second ongoing project (AFTOL-2) (Blackwell et al.,
2006). A major product of the Deep Hypha project was the publication of 24
papers on fungal phylogeny in a single journal issue (Mycologia 98: 829–1103).
The papers included an introduction to progress in fungal phylogeny, a paper on
dating the origin of Fungi, one on the evolution of morphological traits, and 21
articles with multilocus phylogenies of most major groups. Participants included
156 authors with some involved in more than one paper; only 72 of the authors
were originally from North America. The multi-investigator AFTOL-1 publica-
tion (Hibbett et al., 2007) included a widely used and often cited phylogenetic
classification to the level of order (e.g., Kirk et al., 2008; The NCBI Entrez
Taxonomy Home-page, http://www.ncbi.nlm.nih.gov/taxonomy; Science Watch,
http://sciencewatch.com/dr/nhp/2009/09jannhp/09jannhpHibb). The paper in-
cluded 68 authors from more than 20 countries.
It is important to note that there was broad participation and, essentially,
global involvement on these projects, emphasizing that studies of biodiversity are
indeed global endeavors. Additional pages were contributed to the Tree of Life
web project (http://www.tolweb.org/Fungi/2377) to make information on fungi
more accessible to students and the general public. Two objectives of the ongo-
ing AFTOL-2 project include increased taxon sampling of fungi for molecular
144 FUNGAL DISEASES
Figure A3-2-10.eps
bitmap
APPENDIX A 145
FIGURE A3-11╇ Numbers of known fungi from the Dictionary of the Fungi (editions
1–10, 1950–2008). Authors state that the large increase in species numbers in the 10th
edition may be inflated becauseFigure A3-11.eps
asexual and sexual forms were counted separately and
bitmap
molecular techniques that distinguish close taxa have been used.
APPENDIX A 147
forest and three in a mixed tropical forest. Their current collections contain 1200
morphospecies, primarily basidiomycetes. Approximately 260 species were col-
lected repeatedly only in the Dicymbe plots. Thus far, two new genera and ca.
50 new species have been described. On the basis of groups already studied,
Aime estimated that ca. 120 new ectomycorrhizal taxa have been discovered.
Including novel saprobes as well as ectomycorrhizal fungi, ca. 500 new species
are expected among the 1200 taxa collected. It is clear, however, that these are
not simply high numbers of new taxa, but biologically interesting fungi as well
(Aime et al., 2010). One species is so unusual, that a reviewer of the original
report called it “the find of the century” (Redhead, 2002). As Aime has quipped
“if one were to compare the ratio of fungi to plants in the Dicymbe plots as did
Hawksworth (1991), the ratio would be 260 to 1, obviously an overestimate but
also a cautionary exercise in basing any estimate on a single ecotype” (M. C.
Aime, Louisiana State University, personal communication, August 2010).
Many fungi have in fact come from temperate regions, and some studies
report a high diversity of fungi. For example, in a study of indoor air from build-
ings using culture-independent sampling methods, diversity was found to be
significantly higher in temperate sites independent of building design or use. The
authors also alluded to the possibility that previous studies of certain mycorrhizal
fungi showed similar trends (Amend et al., 2010b). More investigation in this
area is needed, but it is clear that many undescribed fungi are present in temper-
ate regions. Popular literature often rationalizes the need to save the rainforests,
not because of their intrinsic value, but because of the potential drug-producing
organisms that may be found there. Many of the commercially most success-
ful fungal drugs, however, come from temperate fungi. Penicillium chrysoge-
num, producer of penicillin, was found in a northern temperate city. Another
remarkable fungus, Tolypocladium infl atum from Norwegian soil, synthesizes
cyclosporine, an immune-suppressant drug that revolutionized organ transplants
(Borel, 2002); the sexual state of this fungus was collected in New York, USA
(Hodge et al., 1996). Today the drug is commonly used to treat dry eye (Perry
et al., 2008), as well as many serious conditions. Statins produced by fungi such
as Aspergillus terreus from temperate regions, combat high cholesterol levels,
as well as providing other benefits (Vaughan et al., 1996; Askenazi et al., 2003;
Baigent et al., 2005).
In temperate deserts, mycorrhizal boletes, agarics, and rust and smut fungi,
are common. A surprising number of wood-decaying basidiomycetes have been
discovered on living and dead desert plants, including cacti and are in the Univer-
sity of Arizona, Robert L. Gilbertson Mycological Herbarium (http:// ag.arizona.
edu/mycoherb/herbholdings). When a noted mycologist moved to Arizona early
in his career, he became excited about the new and unreported fungal diversity
found in the desert. His proposed study of the wood-decaying fungi of the So-
noran Desert was poorly received with a comment that wood-decaying fungi
were not present in the desert (R. L. Gilbertson, University of Arizona, personal
communication, August 1979). The Sonoran Desert, however, has many plants
150 FUNGAL DISEASES
(e.g., cacti, ocotillo, and mesquite and other desert legumes) that are substrates
for polypores and resupinate basidiomycetes (e.g., Gilbertson and Ryvarden,
1986, 1987).
Fungi also grow at low temperatures. An example involves fungal deteriora-
tion of historic huts built between 1901 and 1911 for use by Antarctic explorers
including Robert Scott and Ernest Shackleton, and although there are not large
species numbers, it is important not to overlook this fungal habitat in diversity
studies (Held et al., 2005). Lichens have often been reported to be common in
Arctic and Antarctic regions (Wirtz et al., 2008), and yeasts are active under
frozen conditions in the Antarctic (Vishniac, 2006; Amato et al., 2009). In some
cases, a yeast isolated from the Antarctic (based on 28S rDNA barcoding) also
has been reported from varied habitats, including human infections, the gut of
insects, deep seas, and hydro-carbon seeps (Kurtzman and Fell, 1998; Bass
et al., 2007; personal observation). Although some fungi are specialized for cold
regions, others simply occupy a wide variety of environmental conditions.
Many regions and habitats of the world need to be included in fungal dis-
covery. In general, microscopic fungi and those that cannot be cultured are very
poorly known. Parts of Africa remain to be collected for many, although not all,
fungal groups (Crous et al., 2006). Fungi are important as symbionts, and they
are associated with every major group of organisms, bacteria, plants and green
algae, and animals including insects. Because certain under-studied symbiotic
associations are known to include large numbers of fungi, these are a good place
to search for new taxa. The associated organisms also allow for resampling, a
quick way to obtain data about host specificity. Targeting hosts also is a produc-
tive method for discovering fungal fossils, such as those associated with plants of
the Rhynie Chert (Taylor et al., 2004). Examples of diversity in particular fungal
habitats are reviewed in the following sections.
fungi with about 250 described species in a variety of taxa (Gerdemann, 1968;
Schüssler and Walker, 2011; Wang and Qiu, 2006). Evidence from recent mo-
lecular studies, however, indicates that cryptic species with higher levels of host
specificity than previously realized will increase the number of known AM fungi
(Selosse et al., 2006; Smith and Read, 2008). More than 6000 species, mostly of
mushroom-forming basidiomycetes, form ectomycorrhizae with about 10% of
all plant families. Greater host specificity usually occurs in the ectomycorrhizal
fungus–plant associations than in AM associations (Smith and Read, 2008). Vast
parts of the world remain to be sampled (Mueller et al., 2007), and it is expected
that barriers to inter-breeding have led to high genetic diversity among these fungi
(Petersen and Hughes, 2007).
Plant Pathogens
Plant pathogens differ from endophytes in that they cause disease symptoms.
Although some zoosporic and zygosporic fungi are plant pathogens, most plant
pathogens are ascomycetes and basidiomycetes. A large number of ascomycetes
and ca. 8000 species of basidiomycetes are plant pathogens. In addition to crop
pathogens, it is important to remember that many pathogens are numerous and
important in natural ecosystems (Farr et al., 1989; Burdon, 1993). Nonpathogenic
phylloplane yeasts occupy leaf surfaces of many plants and are increasingly rec-
ognized for their control of potential leaf pathogens (Fonseca and Inácio, 2006).
In addition to the thousands of native fungi that parasitize plants in the United
States, pathologists are constantly on the lookout for introduced pathogens that
often are undescribed when they arrive to decimate naïve native plant popula-
tions. For example, invasive fungi such as those grouped as Dutch elm disease
fungi, chestnut blight fungus, dogwood anthracnose fungus, and redbay wilt
fungus, were all unknown until they were observed soon after their introduction
(Alexopoulos et al., 1996; Zhang and Blackwell, 2001; Harrington et al., 2008).
Exotic localities will need to be searched for undescribed fungi that probably go
largely unnoticed on their native hosts. It is important to note that although fungi
may cause only minor symptoms to hosts in their native habitats, one of these
may have the potential to be the next destructive disease after introduction to a
new region.
Molecular methods have helped to clarify limits of closely related species
and to establish host ranges (e.g., Crous et al., 2008). In a study of 26 leaf spot
fungi in Australia, three genera of Myrtaceae, including Eucalyptus, were hosts
for three new genera and 20 new species (Cheewangkoon et al., 2009). Although
the authors acknowledged the high level of new taxa discovered, they pointed
out that the potential for host shifts within plantations might lower estimates of
fungal species numbers worldwide. Host or substrate specificity is a concept that
can be applied to fungal groups that are closely associated with hosts such as en-
dophytes, pathogens, and mycorrhizal fungi but not usually for saprobic species
(Zhou and Hyde, 2001). In the past species of plant pathogens often were based
on host identity, a practice that is not always effective because some groups are
host-specific while others are not.
Harrington et al., 2008; Aanen et al., 2009). Many yeasts are associated with
insects, particularly insects that feed on nectar (Lachance, 2006; Robert et al.,
2006).
Other insects contain gut yeasts, a habitat where few have looked for them.
Isolations from the gut of mushroom-feeding beetles yielded up to 200 new spe-
cies of yeasts (Suh et al., 2004, 2005; see also Lachance et al., 2010). Because
only about 1500 ascomycete yeasts (Saccharomycotina) have been described, the
gut yeasts represent a dramatic increase in diversity from a limited geographical
range (Boekhout, 2005; C. Kurtzman, USDA-ARS, personal communication,
July 2010). In fact, the estimated total number of yeast species worldwide could
be increased by as much as 50% by simply recollecting in previously collected
sites from the study (Suh et al., 2005). As Lachance (2006) pointed out, based
on predictions of yeast numbers using data from species in slime fluxes and in
associations with flower-visiting insects, it is necessary to obtain more informa-
tion on specificity and geographical ranges before better estimates can be made.
Although not all insects harbor large numbers of yeasts in their guts, those with
restricted diets in all life history stages such as mushrooms or wood are often
associated with yeasts. Host insects may acquire digestive enzymes or vitamins
from the yeasts. This contention is supported by the fact that unrelated insects
feeding on mushrooms (e.g., beetles in different lineages, lepidopteran larvae) all
have gut yeasts with similar assimilative capabilities and vitamin production. The
high rate of discovery of yeasts in under-collected habitats and localities suggests
that far more taxa await discovery (Suh et al., 2005), and the gut habitat has been
considered a yeast diversity hotspot (Boekhout, 2005).
Insects may be food for fungi, especially in low nitrogen environments. The
mycelium of Pleurotus ostreatus, a favorite edible species for humans, secretes
toxic droplets that kill nematodes. A study involving the mushroom-producing,
ectomycorrhizal basidiomycete, Laccaria bicolor, was designed to determine
the amount of predation by springtails on the fungal mycelium. The study led
to the surprise discovery that the fungus was not insect food, but rather, it, and
indirectly, the host tree benefited by obtaining substantial amounts of nitrogen
from the insects (Klironomos and Hart, 2001). The predatory habit has arisen
independently on several occasions in at least four phyla of fungi and oomyce-
tes. Predaceous fungi such as species of Arthrobotrys and Dactylella lure, then
trap, snare, or grip nematodes and other small invertebrate animals in soils and
in wood (Barron, 1977).
Ødegaard (2000) revised global estimates of arthropods downward from 30
million to 5–10 million. Not all insects and arthropods are tightly associated with
fungi, but even the revised species estimates indicate that the numbers of insect-
associated fungi will be very high.
APPENDIX A 155
Soil Fungi
Soil is a habitat of high fungal diversity (Waksman, 1922; Gilman, 1957;
Kirk et al., 2004; Domsch et al., 2007). Soil fungi and bacteria are important
in biogeochemical cycles (Vandenkoornhuyse et al., 2002), and the diversity of
soil fungi is highest near organic material such as roots and root exudates. Per
volume, large numbers of microscopic fungi occur in pure soil, and these are
largely asexual ascomycetes and some zygomycetes, including animal-associated
Zoopagales. Gams (2006) estimated that 3150 species of soil fungi are known,
and ca. 70% are available in culture. There presently is a high rate of new spe-
cies acquisition, and the group appears to be better known than most ecologically
defined groups. Molecular studies, however, are predicted to increase the total
number (Bills et al., 2004). In fact a study of soil communities in several for-
est types at the Bonanza Creek Long Term Ecological Research site, Fairbanks,
Alaska, United States, revealed not only seasonal changes in community compo-
sition but also in dominance of fungi over bacteria. The data acquired by several
molecular methods including high-throughput sequencing greatly increased the
total number of fungal sequences in GenBank at the time (Taylor et al., 2010).
Taylor and his colleagues found more than 200 operational taxonomic units in a
0.25 g soil sample with only 14% overlap in a sample taken a meter away. This
study is not directly comparable with the soil fungi reported by Gams (2006)
because Gams’ figures excluded fungi such as mycorrhizal species.
Another study of soil fungi based on environmental DNA sequences showed
an unexpected distribution of a group of zoosporic fungi, Chytridiomycota. The
chytrids, were found to be the predominate group of fungi in nonvegetated,
high-elevation soils at sites in Nepal and in the United States in Colorado, where
more than 60% of the clone libraries obtained were from chytrids. A phylogenetic
analysis of the sequences compared with those of a broad selection of known
chytrids, indicated that a diverse group of Chytridiomycota representing three
orders was present (Freeman et al., 2009).
Most major fungal lineages are known from cultures and specimens, but
there have been a few surprises even in well-sampled habitats such as soil. Soil
clone group I (SCGI) represents a major lineage of fungi that occurs in temperate
and tropical soils on three continents, but no one has ever seen or isolated any of
the species into culture (Schadt et al., 2003; Porter et al., 2008).
The phylogenetic position of this lineage, perhaps a new phylum, appeared
as a sister group to the clade of Pezizomycotina–Saccharomycotina (Porter et al.,
2008). Other unexpected higher taxonomic level fungal clades have been detected
from environmental DNA sequences (Vandenkoornhuyse et al., 2002; Jumpponen
and Johnson, 2005; Porter et al., 2008). Another lineage detected by environ-
mental sequences was subjected to fluorescent in situ hybridization (FISH). The
outline of a single-celled, flagellated organism was detected (Jones and Richards,
2009), but apparently none of these fungi has been cultured either. Higher-level
156 FUNGAL DISEASES
bacterial taxa have been discovered by environmental sampling, but this is a far
less common occurrence for fungi (Porter et al., 2008).
Fungi form crusts that stabilize desert soils. Crusts usually are made up
of darkly pigmented ascomycetes, lichens, and nitrogen-fixing cyanobacteria
(States and Christensen, 2001). Rock-inhabiting fungi occur in the surface and
subsurface layers of desert rocks. These darkly pigmented ascomycetes are mem-
bers of the classes Dothideomycetes and Arthoniomycetes, but basidiomycetes
and bacteria may occur in the associations (Kuhlman et al., 2006; Ruibal et al.,
2009). Easily cultured asexual ascomycetes and other fungi also occur in desert
soils, and these include an unusual zygomycete, Lobosporangium transversale
(Ranzoni, 1968), known only from three isolations including Sonoran Desert soil.
Yeasts are well known from American deserts in association with cacti and flies
where they detoxify plant metabolites (Starmer et al., 2006).
Freshwater Fungi
Certain fungi are adapted for life in fresh water. More than 3000 species of
ascomycetes are specialized for a saprobic life style in freshwater habitats where
they have enhanced growth and sporulation (Shearer et al., 2007; Kirk et al.,
2008; Shearer and Raja, 2010). The asci are evanescent, and ascospores have ap-
pendages and sticky spore sheaths, that anchor the spores to potential substrates
in the aquatic environment. Conidia have several dispersal strategies, and these
are designated as Ingoldian (Fig. A3-2) and aero-aquatic (Fig. A3-3) conidia.
Ingoldian conidia are sigmoidal, branched, or tetraradiate and attach to plants
and other material in the water. The conidia float on foam that accumulates at the
banks of streams, especially during heavy runoff, and when the bubbles burst, the
spores may be dispersed for great distances from the water and into trees, where
they can be isolated from water-filled tree holes (Bandoni, 1981; Descals and
Moralejo, 2001; Gönczöl and Révay, 2003). Aero-aquatic fungi have multicel-
lular, often tightly helical conidia with air spaces to make them buoyant on the
surface of slower-moving waters (Fisher, 1977).
Other, less obviously modified fungi are present in water, and some of
these are active in degrading leaves in streams after the heavy autumn leaf fall.
A few specialized freshwater basidiomycetes also are known, and several have
branched conidia similar to those of the Ingoldian ascomycetes. Flagellated fungi
occur in aquatic habitats, including Chytridiomycota, Blastocladiomycota, and
Monoblepharomycota (James et al., 2006). Batrachochytrium dendrobatidis, the
recently described amphibian killer, is an aquatic chytrid (Longcore et al., 1999).
Members of Neocallimastigomycota also live in a specialized largely aquatic
environment, the gut of vertebrate herbivores, where they are essential for diges-
tion of cellulosic substrates.
APPENDIX A 157
Marine Fungi
Marine waters provide a habitat for certain specialized fungi (Kohlmeyer
and Volkmann-Kohlmeyer, 1991), and Hyde et al. (1998) estimated that more
than 1500 species of marine fungi occur in a broad array of taxonomic groups.
Many of these fungi are distinct from freshwater aquatic species, and they may be
saprobic on aquatic plant substrates. Some species have characters such as sticky
spore appendages, indicators of specialization for the marine habitat (Kohlmeyer
et al., 2000).
It is interesting that few fungi from early-diverging lineages have been
reported from marine environments, perhaps in part because mycologists study-
ing these groups sampled more often from fresh water habitats. More recently,
an investigation of deep-sea hydrothermal ecosystems revealed not only novel
species of ascomycetes and basidiomycetes, but also what may be a previously
unknown lineage of chytrids (Le Calvez et al., 2009).
Most marine fungi are ascomycetes and basidiomycetes, and these include
ascomycete and basidiomycete yeasts (Nagahama, 2006). Some of the yeasts
degrade hydrocarbon compounds present in natural underwater seeps and spills
(Davies and Westlake, 1979). Certain ascomycetes are specialists on calcareous
substrates including mollusk shells and cnidarian reefs. Even a few mushroom-
forming basidiomycetes are restricted to marine waters (Binder et al., 2006).
Some fungi use other marine invertebrates as hosts (Kim and Harvell, 2004),
including antibiotic producers that live in sponges (Bhadury et al., 2006; Pivkin
et al., 2006; Wang et al., 2008). A wide variety of fungi considered to be terres-
trial also are found in marine environments. Basidiomycete (i.e., Lacazia loboi)
and ascomycete yeasts, and other fungi including Basidiobolus ranarum, may oc-
cur in marine waters where they infect porpoises and other vertebrates (Kurtzman
and Fell, 1998; Murdoch et al., 2008; Morris et al., 2010).
Fungal Species
Currently, molecular methods provide large numbers of characters for use
in phylogenetic species discrimination (e.g., Kohn, 2005; Giraud et al., 2008). In
the past, biologists relied primarily on phenotype for species delimitation, and
most of the formally described species known today were based on morphology.
In addition, mating tests have been used to distinguish so-called biological spe-
cies, especially among heterothallic basidiomycetes (Anderson and Ullrich, 1979;
Petersen, 1995). The ability to mate, however, may be an ancestral character.
For example, Turner et al. (2010) found evidence that fungi have evolved strong
barriers to mating when they have sympatric rather than allopatric distributions.
Distant populations would not have had strong selective pressure against hybrid-
ization, thereby avoiding production of progeny less fit than conspecific progeny
(e.g., Garbelotto et al., 2007; Stireman et al., 2010). This phenomenon, known
as reinforcement, helps to explain how fungi from different continents can mate
158 FUNGAL DISEASES
Conclusions
Until recently, estimates of numbers of fungi did not include results from
large-scale environmental sequencing methods. Newer estimates based on data
acquired from several molecular methods, however, have predicted as many as
5.1 million species of fungi (O’Brien et al., 2005; Taylor et al., 2010). Mycolo-
gists also are beginning to use high-throughput methods to gain insight into ques-
tions including geographical ranges and host and substrate specificity, topics that
have direct bearing on species numbers (Lumbsch et al., 2008). For example,
high-throughput methods have been used to determine the amount of overlap
between species within a given region by comparing soil samples a meter apart
to find only 14% species overlap (Taylor et al., 2010).
A better estimate of fungal numbers also can be speeded by enlisting more
biologists to accomplish the goal. When amphibian populations first were ob-
served to be dwindling and some species were determined to have disappeared
almost 20 yr earlier, a number of causes, all nonfungal, were suggested as the
explanation. The revelation that a chytrid was involved brought to mind that there
were probably fewer than 10 mycologists in the world who could collect, isolate,
culture, and identify the novel flagellated fungus, Batrachochytrium dendrobati-
dis (Longcore et al., 1999). Since that time interest in and publications on chytrids
have increased dramatically (e.g., Freeman et al., 2009; LeCalvez et al., 2009).
The interest in amphibian disease was in part the impetus for a large number of
recent publications on amphibian decline, but amphibian decline also justified
other projects, including training new chytrid systematists in monographic work.
This effort has resulted in the discovery of many new chytrid species and the
description of five new orders between 2008 and 2010. The rise of AIDS and the
APPENDIX A 159
References
Aanen, D. K., H. H.De Fine Licht, A. J. M. Debets, N. G. Kerstes, R. F. Hoekstra, and J. J. Boomsma.
2009. High symbiont relatedness stabilizes mutualistic cooperation in fungus-growing termites.
Science 326:1103–1106.
Aime, M. C., D. L. Largent, T. W. Henkel, and T. J. Baroni. 2010. The Entolomataceae of the Pa-
karaima Mountains of Guyana IV: New species of Calliderma, Paraeccilia and Trichopilus.
Mycologia 102:633–649.
Ainsworth, G. C., and G. R. Bisby. 1943. Dictionary of the Fungi. Imperial Mycological Institute,
Kew, UK.
Alexopoulos,C. J. C. W., Mims, and M. Blackwell.1996. Introductory mycology. Wiley, New York,
New York, USA.
Amato, P., S. M. Doyle, and B. C. Christner. 2009. Macromolecular synthesis by yeasts under frozen
conditions. Environmental Microbiology 11:589–596.
Amend, A. S. K. A. Seifert, and T. D. Bruns. 2010a. Quantifying microbial communities with 454 pyro-
sequencing: Does read abundance count? Molecular Ecology 10.1111/j.1365-294X.2010.04898.x.
Amend, A. S., K. A. Seifert, R.Samson, and T. D. Bruns. 2010b. Indoor fungal composition is geo-
graphically patterned and more diverse in temperate zones than in the tropics. Proceedings of
the National Academy of Sciences, USA 107:13748–13753.
Anderson, J. B., and R. C.Ullrich, 1979. Biological species of Armillaria in North America. Myco-
logia 71:402–414.
Arnold, A. E. 2007. Understanding the diversity of foliar endophytic fungi: Progress, challenges, and
frontiers. Fungal Biology Reviews 21:51–66.
Arnold, A. E., and F. Lutzoni. 2007. Diversity and host range of foliar fungal endophytes: Are tropical
leaves biodiversity hotspots? Ecology 88:541–549.
Arnold, A. E., Z. Maynard, and G. S. Gilbert. 2001. Fungal endophytes in dicotyledonous neotropical
trees: Patterns of abundance and diversity. Mycological Research 105:1502–1507.
Arnold, A. E., Z. Maynard, G. S. Gilbert, P. D. Coley, and T. A. Kursar. 2000. Are tropical fungal
endophytes hyperdiverse? Ecology Letters 3:267–274.
160 FUNGAL DISEASES
Arnold, A. E., L. C. Mejía, D. Kyllo, E. Rojas, Z. Maynard, N.Robbins, and E. A.Herre. 2003. Fun-
gal endophytes limit pathogen damage in leaves of a tropical tree. Proceedings of the National
Academy of Sciences, USA 100:15649–15654.
Askenazi, M., E. M. Driggers, D. A. Holtzman, T. C. Norman, S. Iverson, D. P. Zimmer, M. E.
Boers, et al. 2003. Integrating transcriptional and metabolite profiles to direct the engineering
of lovastatin-producing fungal strains. Nature Biotechnology 21:150–156.
Baigent, C., A. Keech, P. M. Kearney, L. Blackwell, G. Buck, C. Pollicino, A. Kirby, et al. 2005.
Efficacy and safety of cholesterol-lowering treatment: Prospective meta-analysis of data from
90,056 participants in 14 randomised trials of statins. Lancet 366:1267–1278.
Bandoni, R. J. 1981. Aquatic hyphomycetes from terrestrial litter. In D. T. Wicklow and G. C. Carroll
[eds.], The fungal community: Its organization and role in the ecosystem, 693–708. Marcel
Dekker, New York, New York, USA.
Barron, G. L. 1977. The nematode destroying fungi. Canadian Biological Publishers, Guelph, Ontario,
Canada.
Bass, D., A. Howe, N. Brown, H. Barton, M. DeMidova, H. Michelle, L. Li, et al. 2007. Yeast forms
dominate fungal diversity in the deep oceans. Proceedings of the Royal Soceity of London, B,
Biological Sciences 274:3069–3077.
Begerow, D., H. Nilsson, M. Unterseher, and W. Maier, 2010. Current state and perspectives of fungal
DNA barcoding and rapid identification procedures. Applied Microbiology and Biotechnology
87:99–108.
Benjamin, R. K., M. Blackwell, I. Chapella, R. A. Humber, K. G. Jones, K. A. Klepzig, R. W. Lich-
twardt, et al. 2004. The search for diversity of insects and other arthropod associated fungi. In G.
M. Mueller, G. F. Bills, and M. S. Foster [eds.], Biodiversity of fungi: Inventory and monitoring
methods, 395–433. Elsevier Academic Press, San Diego, California, USA.
Berbee, M. L., and J. W. Taylor. 2010. Dating the molecular clock in fungi—How close are we?
Fungal Biology Reviews 24:1–16.
Bhadury, P., B. T. Mohammad, and P. C. Wright. 2006. The current status of natural products from
marine fungi and their potential as anti-infective agents. Journal of Industrial Microbiology &
Biotechnology 33:325–337.
Bills, G. F., M. Christensen, M. J. Powell, and G. Thorn. 2004. Saprobic soil fungi. In G. M. Mueller,
G. F. Bills, and M. S. Foster [eds.], Biodiversity of fungi: Inventory and monitoring methods,
271–302. Elsevier Academic Press, San Diego, California, USA.
Binder, M., D. S. Hibbett, Z. Wang, and W. F. Farnham. 2006. Evolutionary relationships of My-
caureola dilseae (Agaricales), a basidiomycete pathogen of a subtidal rhodophyte. American
Journal of Botany 93:547–556.
Bisby, G. R., and G. C. Ainsworth. 1943. The numbers of fungi. Transactions of the British Myco-
logical Society 26:16–19.
Blackwell, M., D. S. Hibbett, J. W. Taylor, and J. W. Spatafora. 2006. Research coordination net-
works: A phylogeny for kingdom Fungi (Deep Hypha). Mycologia 98:829–837.
Boekhout, T. 2005. Gut feeling for yeasts. Nature 434:449–451.
Borel, J. F. 2002. History of the discovery of cyclosporin and of its early pharmacological develop-
ment. Wiener Klinische Wochenschrift 114:433–437.
Bowman, B. H., J. W. Taylor, A. G. Brownlee, J. Lee, S.-D. Lu, and T. J. White. 1992. Molecular
evolution of the fungi: Relationship of the Basidiomycetes, Ascomycetes and Chytridiomycetes.
Molecular Biology and Evolution 9:285–296.
Brock, P. M., H. Doring, and M. I. Bidartondo. 2009. How to know unknown fungi: The role of a
herbarium. New Phytologist 181:719–724.
Burdon, J. J. 1993. The structure of pathogen populations in natural plant communities. Annual Re-
view of Phytopathology 31:305–323.
Cafaro, M. J. 2005. Eccrinales (Trichomycetes) are not fungi, but a clade of protists at the early
divergence of animals and fungi. Molecular Phylogenetics and Evolution 35:21–34.
APPENDIX A 161
Carroll, G. C. 1988. Fungal endophytes in stems and leaves: From latent pathogen to mutualistic
symbiont. Ecology 69:2–9.
Cavalier-Smith, T. 1983. A 6-kingdom classification and a unified phylogeny. In H. E. A. Chenk and
W. S. Schwemmler [eds.], Endocytobiology II: Intracellular space as oligogenetic, 1027–1034.
Walter de Gruyter, Berlin, Germany.
Celio, G. J., M. Padamsee, B. T. Dentinger, R. Bauer, and D. J. McLaughlin. 2006. Assembling
the Fungal Tree of Life: Constructing the structural and biochemical database. Mycologia
98:850–859.
Cheewangkoon, R., J. Z. Groenwald, B. A. Summerell, K. D. Hyde, C. To-Anun, and P. W. Crous.
2009. Myrtaceae, a cache of fungal biodiversity. Persoonia 23:55–85.
Clay, K., S. Marks, and G. P. Cheplick. 1993. Effects of insect herbivory and fungal endophyte infec-
tion on competitive interactions among grasses. Ecology 74:1767–1777.
Corradi, N., and P. J. Keeling. 2009. Microsporidia: A journey through radical taxonomic revisions.
Fungal Biology Reviews 23:1–8.
Crous, P. W., I. H. Rong, A. Wood, S. Lee, H. Glen, W. Botha, B. Slippers, et al. 2006. How many
species of fungi are there at the tip of Africa? Studies in Mycology 55:13–33.
Crous, P. W., B. A. Summerell, L. Mostert, and J. Z. Groenewald. 2008. Host specificity and specia-
tion of Mycosphaerella and Teratosphaeria species associated with leaf spots of Proteaceae.
Persoonia 20:59–86.
Davies, J. S., and D. W. S. Westlake. 1979. Crude oil utilization by fungi. Canadian Journal of
Microbiology 25:146–156.
De Kesel, A. 1996. Host specificity and habitat preference of Laboulbenia slackensis. Mycologia
88:565–573.
Descals, E., and E. Moralejo. 2001. Water and asexual reproduction in the Ingoldian fungi. Botanica
Complutensis 25:13–71.
Dettman, J. R., D. J. Jacobson, and J. W. Taylor. 2006. Multilocus sequence data reveal extensive
phylogenetic species diversity within the Neurospora discreta complex. Mycologia 98:436–446.
Diehl, W. W. 1950. Balansia and the Balansiae in America. USDA Agriculture Monograph 4:1–82.
Domsch, K. H., W. Gams, and T. H. Anderson. 2007. Compendium of soil fungi, 2nd ed. IHW-Verlag
and Verlagsbuchhandlung, Eching, Germany.
Farr, D. F., G. F. Bills, G. P. Chamuris, and A. Y. Rossman. 1989. Fungi on plants and plant products in
the United States, 2nd ed. American Phytopathological Society Press, St. Paul, Minnesota, USA.
Feuerer, T. [ed.]. 2010. The index of checklists of lichens and lichenicolous fungi [online]. Web-
site http://www.biologie.uni-hamburg.de/checklists/lichens/portalpages/portalpage_checklists_
switch.htm [accessed 30 January 2011].
Feuerer, T., and D. L. Hawksworth. 2007. Biodiversity of lichens, including a world-wide analysis of
checklist data based on Takhtajan’s floristic regions. Biodiversity and Conservation 16:85–98.
Fisher, M. C., G. L. Koenig, T. J. White, and J. W. Taylor. 2002. Molecular and phenotypic description
of Coccidioides posadasii sp. nov., previously recognized as the non-California population of
Coccidioides immitis. Mycologia 94:73–84.
Fisher, P. J. 1977. New methods of detecting and studying saprophytic behaviour of aero-aquatic
hyphomycetes. Transactions of the British Mycological Society 68:407–411.
Fonseca, Á., and J. Inácio. 2006. Phylloplane yeasts. In C. Rosa and P. Gábor [eds.], Biodiversity and
ecophysiology of yeasts, 63–301. Springer-Verlag, Berlin, Germany.
Freeman, K. R., A. P. Martin, D. Karki, R. C. Lynch, M. S. Mitter, A. F. Meyer, J. E. Longcore, et al.
2009. Evidence that chytrids dominate fungal communities in high-elevation soils. Proceedings
of the National Academy of Sciences, USA 106:18315–18320.
Gams, W. 2006. Biodiversity of soil-inhabiting fungi. Biodiversity and Conservation 16:69–72.
Garbelotto, M., P. Gonthier, and G. Nicolotti. 2007. Ecological constraints limit the fitness of fungal
hybrids in the Heterobasidion annosum species complex. Applied and Environmental Micro-
biology 73:6106–6111.
162 FUNGAL DISEASES
Gerdemann, J. W. 1968. Vesicular arbuscular mycorrhiza and plant growth. Annual Review of Phy-
topathology 6:397–418.
Gilbertson, R. L., and M. Blackwell. 1984. Two new basidiomycetes on living live oak in the south-
east and Gulf Coast region. Mycotaxon 20:85–93.
Gilbertson, R. L., and L. Ryvarden. 1986. North American polypores, vol. I. Abortiporus-Lindtneria.
Fungiflora Press, Oslo, Norway.
Gilbertson, R. L., and L. Ryvarden. 1987. North American polypores, vol. II. Megasporoporia-
Wrightoporia. Fungiflora Press, Oslo, Norway.
Gilman, J. C. 1957. A manual of soil fungi, 2nd ed. Iowa State College Press, Ames, Iowa, USA.
Giraud, T., G. Refrégier, M. Le Gac, D. M. De Vienne, AND M. E. Hood. 2008. Speciation in fungi.
Fungal Genetics and Biology 45:791–802.
Gönczöl, J., and Á. Révay. 2003. Treehole fungal communities: Aquatic, aero-aquatic and dematia-
ceous hyphomycetes. Fungal Diversity 12:19–24.
Harrington, T. C., S. W. Fraedrich, and D. N. Aghayeva. 2008. Raffaelea lauricola, a new ambrosia
beetle symbiont and pathogen on the Lauraceae. Mycotaxon 104:399–404.
Hawksworth, D. L. 1991. The fungal dimension of biodiversity: Magnitude, significance, and con-
servation. Mycological Research 95:641–655.
Hawksworth, D. L. 2001. The magnitude of fungal diversity: The 1.5 million species estimate revis-
ited. Mycological Research 105:1422–1432.
Hawksworth, D. L., and A. Y. Rossman. 1997. Where are all the undescribed fungi? Phytopathology
87:888–891.
Held, B. W., J. A. Jurgens, B. E. Arenz, S. M. Duncan, R. L. Farrell, and R. A. Blanchette. 2005.
Environmetal factors influencing microbial growth inside the historic huts of Ross Island, Ant-
arctica. International Biodeterioration & Biodegradation 55:45–53.
Hibbett, D. M., M. Binder, J. F. Bischoff, M. Blackwell, P. F. Cannon, O. Eriksson, S. Huhndorf,
et al. 2007. A higher-level phylogenetic classification of the Fungi. Mycological Research
111:509–547.
Hibbett, D. S., A. Ohman, D. Glotzer, M. Nuhn, P. Kirk, and R. H. Nilsson. In press. Progress in
molecular and morphological taxon discovery in Fungi and options for formal classification of
environmental sequences. Fungal Biology Reviews.
Hillebrand, H. 2004. On the generality of the latitudinal diversity gradient. American Naturalist
163:192–211.
Hodge, K. T., S. B. Krasnoff, and R. A. Humber. 1996. Tolypocladiuminfl atum is the anamorph of
Cordyceps subsessilis. Mycologia 88:715–719.
Hyde, K. D. 2001. Where are the missing fungi? Mycological Research 105:1409–1412.
Hyde, K. D., E. B. G. Jones, E. Leaño, S. B. Pointing, A. D. Poonyth, and L. L. P. Vrijmoed. 1998.
Role of fungi in marine ecosystems. Biodiversity and Conservation 7:1147–1161.
Innis, M. A., D. H. Gelfand, J. J. Sninsky, and T. J. White. 1990. PCR protocols: A guide to methods
and applications. Academic Press, San Diego, California, USA.
Jacobson, D. J., J. R. Dettman, R. I. Adams, C. Boesl, S. Sultana, T. Roenneberg, M. Merrow, et al.
2006. New findings of Neurospora in Europe and comparisons of diversity in temperate climates
on continental scales. Mycologia 98:550–559.
James, T. Y., P. M. Letcher, J. E. Longcore, S. E. Mozley-Standridge, D. Porter, M. J. Powell, G. W.
Griffith, and R. Vilgalys. 2006. A molecular phylogeny of the flagellated Fungi (Chytridiomy-
cota) and description of a new phylum (Blastocladiomycota). Mycologia 98:860–871.
Jones, M. D. M., and T. A. Richards. 2009. Environmental DNA combined with fluorescent in situ
hybridisation reveals a missing link in the fungal tree of life. Proceedings of 25th Fungal Genet-
ics Conference, 2009, Asilomar, California, USA, abstract 427.
Joppa, L. N., D. L. Roberts, and S. L. Pimm. 2010. How many species of flowering plants are there?
Proceedings of the Royal Society of London, B, Biological Sciences 278:554–559.
Jumpponen, A., and L. C. Johnson. 2005. Can rDNA analyses of diverse fungal communities in soil
and roots detect effects of environmental manipulations—A case study from tallgrass prairie.
Mycologia 97:1177–1194.
APPENDIX A 163
Jumpponen, A., and K. L. Jones. 2009. Massively parallel 454 sequencing indicates hyperdiverse fun-
gal communities in temperate Quercus macrocarpa phyllosphere. New Phytologist 184:438–448.
Keeling, P. J. 2009. Five questions about Microsporidia. PLoS Pathogens 5: e1000489.
Kim, K., and C. D. Harvell. 2004. The rise and fall of a six year coral fungal epizootic. American
Naturalist 164:S52–S63.
Kirk, J. L., L. A. Beaudette, M. Hart, P. Moutoglis, J. N. Klironomos, H. Lee, and J. T. Trevors. 2004.
Methods of studying soil microbial diversity. Journal of Microbiological Methods 58:169–188.
Kirk, P. M., P. F. Cannon, D. W. Minter, and J. A. Stalpers. 2008. Dictionary of the Fungi, 10th ed.
CABI, Wallingford, UK.
Klironomos, J. N., and M. M. Hart. 2001. Animal nitrogen swap for plant carbon. Nature 410:651–652.
Kohlmeyer, J., J. W. Spatafora, and B. Volkmann-Kohlmeyer. 2000. Lulworthiales, a new order of
marine Ascomycota. Mycologia 92:453–458.
Kohlmeyer, J., and B. Volkmann-Kohlmeyer. 1991. Illustrated key to the filamentous higher marine
fungi. Botanica Marina 34:1–61. Kohn, L. M. 2005. Mechanisms of fungal speciation. Annual
Review of Phytopathology 43:279–308.
Kuhlman,K. R.,W. G.Fusco,M. T.La Duc,L. B.Allenbach,C. L.Ball, G. M. Kuhlman, R. C. Anderson,
et al. 2006. Diversity of microorganisms within rock varnish in the Whipple Mountains, Cali-
fornia. Applied and Environmental Microbiology 72:1708–1715.
Kurtzman,C. P., and J. W. Fell. 1998. The yeasts, a taxonomic study, 4th ed. Elsevier, Amsterdam,
Netherlands.
Kurtzman, C. P., and C. J. Robnett. 1998. Identification and phylogeny of ascomycetous yeasts from
analysis of nuclear large subunit (26S) ribosomal DNA partial sequences. Antonie van Leeu-
wenhoek 73:331–371.
LaChance, M.-A. 2006. Yeast biodiversity: How many and how much? In C. Rosa and P. Gábor [eds.],
Biodiversity and ecophysiology of yeasts, 1–9. Springer-Verlag, Berlin, Germany.
LaChance, M.-A., J. Dobson, D. N. Wijayanayaka, and A. M. E. Smith. 2010. The use of parsimony
network analysis for the formal delineation of phylogenetic species of yeasts: Candida apicola,
Candida azyma, and Candida parazyma sp. nov., cosmopolitan yeasts associated with florico-
lous insects. Antonie van Leeuwenhoek 97:155–170.
Le Calvez, T., G. Burgaud, S. Mahé, G. Barbier, and P. Vandenkoornhuyse. 2009. Fungal diversity
in deepsea hydrothermal ecosystems. Applied and Environmental Microbiology 75:6415–6421.
Lee, S. C., N. Corradi, S. Doan, F. S. Dietrich, P. J. Keeling, and J. Heitman. 2010. Evolution of the
sex-related locus and genomic features shared in Microsporidia and Fungi. PLoS ONE 5:e10539.
10.1371/journal.pone.0010539.
Lichtwardt, R. W., M. J. Cafaro, and M. M. White. 2001 The Trichomycetes: Fungal associates of
arthropods, revised ed. [online]. Website http://www.nhm.ku.edu/~fungi [accessed 30 January
2011].
Liti, G., D. M. Carter, A. M. Moses, J. Warringer, L. Parts, S. A. James, R. P. Davey, et al. 2009.
Population genomics of domestic and wild yeasts. Nature 458:337–341.
Little, A. E. F., and C. R. Currie. 2007. Symbiont complexity: Discovery of a fifth symbiont in the
attine ant–microbe symbiosis. Biology Letters 3:501–504.
Longcore, J. E., A. P. Pessier, and D. K. Nichols. 1999. Batrachochytrium dendrobatidis gen. et sp.
nov., a chytrid pathogenic to amphibians. Mycologia 91:219–227.
Lumbsch, H. T., P. K. Buchanan, T. W. May, and G. M. Mueller. 2008. Phylogeography and bioge-
ography of Fungi. Mycological Research 112:423–484.
Lutzoni, F., and J. Miadlikowska. 2009. Lichens. Current Biology 19:R502–R503.
Martin, G. W. 1951. The numbers of fungi. Proceedings of the Iowa Academy of Science 58:175–178.
Medina, M., A. G. Collins, J. W. Taylor, J. W. Valentine, J. H. Lipps, L. A. Amaral-Zettler, and M. L.
Sogin. 2003. Phylogeny of Opistokonta and the evolution of multicellularity and complexity in
Fungi and Metazoa. International Journal of Astrobiology 2:203–211.
Metzker, M. L. 2010. Sequencing technologies—The next generation. Nature Reviews Genetics
11:31–46.
164 FUNGAL DISEASES
Morris, P. J., W. R. Johnson, J. Pisanic, G. D. Bossart, J. Adams, J. S. Reif, and P. A. Fair. 2010. Isola-
tion of culturable microorganisms from free-ranging bottle nose dolphins (Tursiops truncatus)
from the southeastern United States. Veterinary Microbiology 10.1016/j. vetmic.2010.08.025.
Mueller, G. M., and J. P. Schmit. 2007. Fungal biodiversity: What do we know? What can we predict?
Biodiversity and Conservation 16:1–5.
Mueller, G. M., J. P.Schmit, P. R.Leacock, B. Buyck, J. Cifuentes, D. E. DesJardin, R. E. Halling,
et al. 2007. Global diversity and distribution of macrofungi. Biodiversity and Conservation
16:37–48.
Murdoch, M. E., J. S. Reif, M. Mazzoil, S. D. McCulloch, P. A. Fair, and G. D. Bossart. 2008. Lo-
bomycosis in bottlenose dolphins (Tursiops truncatus) from the Indian River Lagoon, Florida:
Estimation of prevalence, temporal trends, and spatial distribution. EcoHealth 5:289–297.
Nagahama, T. 2006. Yeast biodiversity in freshwater, marine and deep-sea environments. In C. Rosa
and P. Gábor [eds.], Biodiversity and ecophysiology of yeasts, 241–262. Springer-Verlag, Berlin,
Germany.
Nash, T. H. III, B. D. Ryan, P. Diederich, C. Gries, and F. Bungartz. 2004. Lichen flora of the greater
Sonoran Desert region, vol. 2, Most of the microlichens, balance of the macrolichens, and the
lichenicolous fungi. Lichen Unlimited, Tempe, Arizona, USA.
Nash, T. H. III, B. D. Ryan, C. Gries, and F. Bungartz.2002. Lichenflora of the greater Sonoran
Desert region, vol. 1, The pyrenolichens and most of the squamulose and marolichens. Lichen
Unlimited, Tempe, Arizona, USA.
Neafsey, D. E., B. M. Barker, T. J. Sharpton, J. E. Stajich, D. J. Park, E. Whiston, C.-Y. Hung, et al.
2010. Population genomic sequencing of Coccidioides fungi reveals recent hybridization and
transposon control. Genome Research 20:938–946.
Nilsson, R. H., M. Ryberg, E. Kristiansson, K. Abarenkov, K.-H. Larsson, and U. Kõljalg. 2006.
Taxonomic reliability of DNA sequences in public sequence databases: A fungal perspective.
PLoS ONE 1:e59. 10.1371/journal.pone.0000059.
O’Brien, B. L., J. L. Parrent, J. A. Jackson, J. M. Moncalvo, and R. Vilgalys. 2005. Fungal commu-
nity analysis by large-scale sequencing of enviromental samples. Applied and Environmental
Microbiology 71:5544–5550.
Ødegaard, F. 2000. How many species of arthropods? Erwin’s estimate revised. Biological Journal
of the Linnean Society 71:583–597.
Otrosina, W. J., and M. Garbelotto. 2010. Heterobasidion occidentale sp. nov. and Heterobasidion
irregulare nom. nov.: A disposition of North American Heterobasidion biological species.
Fungal Biology 114:16–25.
Paton, A. J., N. Brummitt, R. Govaerts, K. Harman, S. Hinchcliffe, B. Allkin, and E. N. Lughadha.
2008. Towards Target 1 of the Global Strategy for Plant Conservation: A working list of all
known plant species—Progress and prospects. Taxon 57:602–611.
Penfound, W. T., and F. P. Mackaness. 1940. A note concerning the relation between drainage pattern,
bark conditions, and the distribution of corticolous bryophytes. Bryologist 43:168–170.
Perry, H. D., R. Solomon, E. D. Donnenfeld, A. R. Perry, J. R. WittpenN, H. E. Greenman, and H. E.
Savage. 2008. Evaluation of topical cyclosporine for the treatment of dry eye disease. Archives
of Ophthalmology 126:1046–1050.
Petersen, R. H. 1995. There’s more to a mushroom than meets the eye: Mating studies in the Agari-
cales. Mycologia 87:1–17.
Petersen, R. H., and K. W. Hughes. 2007. Some agaric distributions involving Pacific landmasses and
Pacific Rim. Mycoscience 48:1–14.
Pianka, E. R. 1966. Latitudinal gradients in species diversity: A review of concepts. American Natu-
ralist 100:33–46.
Pinruan, U., K. D. Hyde, S. Lumyong, E. H. C. McKenzie, and E. B. G. Jones. 2007. Occurrence of
fungi on tissues of the peat swamp palm Licuala longicalycata. Fungal Diversity 25:157–173.
APPENDIX A 165
Pivkin, M. V., S. A. Aleshko, V. B. Krasokhin, and YU. V.Khudyakova. 2006. Fungal assemblages
associated with sponges of the southern coast of Sakhalin Island. Russian Journal of Marine
Biology 32:207–213.
Porter, T. M., C. W. Schadt, L. Rizvi, A. P. Martin, S. K. Schmidt, L. Scott-Denton, R. Vilgalys,
and J. M. Moncalvo. 2008. Widespread occurrence and phylogenetic placement of a soil clone
group adds a prominent new branch to the fungal tree of life. Molecular Phylogenetics and
Evolution 46:635–644.
Pressel, S., M. I. Bidartondo, R. Ligrone, and J. G. Duckett. 2010. Fungal symbioses in bryophytes:
New insights in the twenty first century. Phytotaxa 9:238–253.
Pringle, A., J. D. Bever, M. Gardes, J. L. Parrent, M. C. Rillig, and J. N. Klironomos. 2009. Mycor-
rhizal symbioses and plant invasions. Annual Review of Ecology, Evolution, and Systematics
40:699–715.
Ranzoni, F. V. 1968. Fungi isolated in culture from soils of the Sonoran Desert. Mycologia 60:356–371.
Raspor, P., and J. Zupan. 2006. Yeasts in extreme environments. In C. Rosa and P. Gábor [eds.], Bio-
diversity and ecophysiology of yeasts, 372–417. Springer-Verlag, Berlin, Germany.
Redhead, S. 2002. Pseudotulostoma: The find of the century? Inoculum 53:2.
Robert, V., J. Stalpers, T. Boekhout, and S.-H. Tan. 2006. Yeast biodiversity and culture collections.
In C. Rosa and P. Gábor [eds.], Biodiversity and ecophysiology of yeasts, 31–44. Springer-
Verlag, Berlin, Germany.
Rodriguez, R. J., J. F. White JR., A. E. Arnold, and R. S. Redman. 2009. Fungal endophytes: Diversity
and functional roles. New Phytologist 182:314–330.
Rossi, W., and A. Weir. 2007. New species of Corethromyces from South America. Mycologia
99:131–134.
Rossman, A. 1994. A strategy for an all-taxa inventory of fungal biodiversity. In C. I. Peng and C.
H. Chou [eds.], Biodiversity and terrestrial ecosystems, 169–194. Academia Sinica Monograph
Series no. 14, Taipei, Taiwan.
Ruibal, C., C. Gueidan, L. Selbmann, A. A. Gorbushina, P. W. Crous, J. Z. Groenewald, L. Muggia,
et al.. 2009. Phylogeny of rock inhabiting fungi related to Dothideomycetes. Studies in Mycol-
ogy 64:123–133.
Saikkonen, K., S. H. Faeth, M. Helander, and T. J. Sullivan. 1998. Fungal endophytes: A continuum
of interactions with host plants. Annual Review of Ecology and Systematics 29:319–343.
Schadt, C. W., A. P. Martin, D. A. Lipson, and S. K. Schmidt. 2003. Seasonal dynamics of previously
unknown fungal lineages in tundra soils. Science 301:1359–1361.
Schmit, J. P., and G. M. Mueller. 2007. An estimate of the lower limit of global fungal diversity.
Biodiversity and Conservation 16:99–111.
Schüssler, A., and C. Walker. 2010. Glomeromycota species list [online]. Website http://www.lrz.
de/~schuessler/amphylo/amphylo_ species.html [accessed 30 January 2011].
Selosse, M. A., F. Richard, X. He, and S. W. Simard. 2006. Mycorrhizal networks: Des liaisons dan-
gereuses? Trends in Ecology & Evolution 21:621–628.
Shearer, C. A., E. Descals, B. Kohlmeyer, J. Kohlmeyer, L. Marvanová, D. Padgett, D. Porter, et al.
2007. Fungal diversity in aquatic habitats. Biodiversity and Conservation 16:49–67.
Shearer, C. A., and H. A. Raja. 2010. Freshwater ascomycetes database [online]. Website http://fungi.
life.illinois.edu/ [accessed 30 January 2011].
Smith, S. E., and D. J. Read. 2008. Mycorrhizal symbiosis, 3rd ed. Academic Press, San Diego,
California, USA.
Spatafora, J. W., G.-H. Sung, and R. Kepler. 2010. An electronic monograph of Cordyceps and related
fungi [online]. Website http:// Cordyceps.us [accessed 30 January 2011].
Spatafora, J. W., G.-H. Sung, J.-M. Sung, N. Hywel-Jones, and J. F.White. 2007. Phylogenetic
evidence for an animal pathogen origin of ergot and the grass endophytes. Molecular Ecology
16:1701–1711.
Stajich, J. E., M. L. Berbee, M. Blackwell, D. S. Hibbett, T. Y. James, J. W. Spatafora, and J. W.
Taylor. 2009. The Fungi. Current Biology 19:R840–R845.
166 FUNGAL DISEASES
Starmer, W. T., V. Aberdeen, and M.-A. LaChance. 2006. The biogeographic diversity of cactophilic
yeasts. In C. Rosa and P. Gábor [eds.], Biodiversity and ecophysiology of yeasts, 486–499.
Springer-Verlag, Berlin, Germany.
States, J. S., and M. Christensen. 2001. Fungi associated with biological soil crusts in desert grass-
lands of Utah and Wyoming. Mycologia 93:432–439.
Stireman, J. O. III, H. P. Devlin, T. G. Carr, and P. Abbot. 2010. Evolutionary diversification of the
gall midge genus Asteromyia (Cecidomyiidae) in a multitrophic ecological context. Molecular
Phylogenetics and Evolution 54:194–210.
Suh, S.-O., J. V. McHugh, and M. Blackwell. 2004. Expansion of the Candida tanzawaensis yeast
clade: 16 novel Candida species from basidiocarp-feeding beetles. International Journal of
Systematic and Evolutionary Microbiology 54:2409–2429.
Suh, S.-O., J. V. McHugh, D. Pollock, and M. Blackwell. 2005. The beetle gut: A hyperdiverse source
of novel yeasts. Mycological Research 109:261–265.
Taylor, D. L., I. C. Herriott, K. E. Stone, J. W. McFarland, M. G. Booth, and M. B. Leigh. 2010.
Structure and resilience of fungal communities in Alaskan boreal forest soils. Canadian Journal
of Forest Research 40:1288–1301.
Taylor, J. W., D. J. Jacobson, S. Kroken, T. Kasuga, D. M. Geiser, D. S. Hibbett, and M. C. Fisher.
2000. Phylogenetic species recognition and species concepts in fungi. Fungal Genetics and
Biology 31:21–32.
Taylor, T. N., S. D. Klavins, M. Krings, E. L. Taylor, H. Kerp, and H. Hass. 2004. Fungi from the
Rhynie Chert: A view from the dark side. Transactions of the Royal Society of Edinburgh, Earth
Sciences 94:457–473.
Tedersoo, L., R. H. Nilsson, K. Abarenkov, T. Jairus, A. Sadam, I. Saar, M. Bahram, et al. 2010. 454
pyrosequencing and Sanger sequencing of tropical mycorrhizal fungi provide similar results but
reveal substantial methodological biases. The New Phytologist 166:1063–1068.
Trappe, J. M. 1987. Phylogenetic and ecologic aspects of mycotrophy in the angiosperms from an
evolutionary standpoint. In G. R. Safir [ed.], Ecophysiology of VA mycorrhizal plants, 2–25.
CRC Press, Boca Raton, Florida, USA.
Turner, E., D. J. Jacobson, and J. W. Taylor. 2010. Reinforced post-mating reproductive isola-
tion barriers in Neurospora, an ascomycete microfungus. Journal of Evolutionary Biology
23:1642–1656.
Vandenkoornhuyse, P., S. L. Baldauf, C. Leyval, J. Straczek, and J. P. W. Young. 2002. Extensive
fungal biodiversity in plant roots. Science 295:2051.
Vaughan, C. J., M. B. Murphy, and B. M. Buckley. 1996. Statins do more than just lower cholesterol.
Lancet 348:1079–1082.
Vega, F. E., A. Simpkins, M. C. Aime, F. Posada, S. W. Peterson, S. A. Rehner, F. Infante, et al. 2010.
Fungal endophyte diversity in coffee plants from Colombia, Hawai’i, Mexico, and Puerto Rico.
Fungal Ecology 3:122–138.
Villalta, C. F., D. J. Jacobson, and J. W. Taylor. 2009. Three new phylogenetic and biological Neuros-
pora species: N. hispaniola, N. metzenbergii and N. perkinsii. Mycologia 101:777–789.
Vishniac, H. S. 2006. Yeast biodiversity in the Antarctic. In C. Rosa and P. Gábor [eds.], Biodiversity
and ecophysiology of yeasts, 419–440. Springer-Verlag, Berlin, Germany.
Vossbrinck, C. R., J. V. Maddox, S. Friedman, B. A. DeBrunner-Vossbrinck, and C. R. Woese. 1987.
Ribosomal RNA sequence suggests microsporidia are extremely ancient eukaryotes. Nature 326:
411–414.
Waksman, S. A. 1922. A method for counting the number of fungi in the soil. Journal of Bacteriol-
ogy 7:339–341.
Wang, B., and Y.-L. Qiu. 2006. Phylogenetic distribution and evolution of mycorrhizas in land plants.
Mycorrhiza 16:299–363.
Wang, G., Q. Li, and P. Zhu. 2008. Phylogenetic diversity of culturable fungi associated with the Ha-
waiian sponges Suberites zeteki and Gelliodes fi brosa. Antonie van Leeuwenhoek 93:163–174.
APPENDIX A 167
Weir, A., and M. Blackwell. 2005. Phylogeny of arthropod ectoparasitic ascomycetes. In F. E. Vega
and M. Blackwell [eds.], Insect–fungal associations: Ecology and evolution, 119–145. Oxford
University Press, New York, New York, USA.
Weir, A., and P. M. Hammond. 1997a. Laboulbeniales on beetles: Host utilization patterns and species
richness of the parasites. Biodiversity and Conservation 6:701–719.
Weir, A., and P. M. Hammond. 1997b. A preliminary assessment of speciesrichness patterns of
tropical, beetle-associated Laboulbeniales (Ascomycetes). In K. D. Hyde [ed.], Biodiversity of
tropical microfungi, 121–139. Hong Kong University Press, Hong Kong.
White, M. M., T. Y. James, K. O’Donnell, M. J. Cafaro, Y. Tanabe, and J. Sugiyama. 2006. Phylogeny
of the Zygomycota based on nuclear ribosomal sequence data. Mycologia 98:872–884.
White, T. J., T. D. Bruns, S. B. Lee, and J. W. Taylor. 1990. Amplification and direct sequencing of
fungal ribosomal RNA Genes for phylogenetics. In M. A. Innis, D. H. Gelfand, J. J. Sninsky, and
T. J. White [eds.], PCR protocols and applications—A laboratory manual, 315–322. Academic
Press, New York, New York, USA.
Wirtz, N., C. Printzen, and H. T. Lumbsch. 2008. The delimitation of Antarctic and bipolar species of
neuropogonoid Usnea (Ascomycota, Lecanorales): A cohesion approach of species recognition
for the Usnea perpusilla complex. Mycological Research 112:472–484.
Zhang, N., and M. Blackwell. 2001. Molecular phylogeny of dogwood anthracnose fungus (Discula
destructiva) and the Diaporthales. Mycologia 93:356–364.
Zhou, D., and K. D. Hyde. 2001. Host-specificity, host-exclusivity, and host-recurrence in saprobic
fungi. Mycological Research 105:1449–1457.
A4
Summary
• The newly described fungus, Geomyces destructans, causes an invasive
skin infection in bats and is the likely agent of white-nose syndrome
(WNS).
16╛╛Reprinted with permission from the American Society for Microbiology (Microbe, June 2011,
pp. 267–273).
17╛╛David S. Blehert is the head of the diagnostic microbiology laboratory at the U.S. Geological
the wing (Figure A4-3A). Gross damage to wing membranes such as depigmenta-
tion, holes, and tears are suggestive of WNS, but these lesions are nonspecific,
and histopathologic examination is necessary to diagnose the disease.
Specifically, fungal invasion of wing membranes ranges from characteristic
cup-like epidermal erosions filled with fungal hyphae to ulceration and invasion
of underlying connective tissue, with fungal invasion sometimes spanning the full
thickness of the wing membrane (Figure A4-3B). Fungal hyphae can also fill hair
follicles and destroy skin glands and local connective tissue. Bat wings play an
important role in the pathogenesis of WNS by providing a large surface area for
the fungus to colonize. Once infected, the thin layer of skin that composes the
bat wing is vulnerable to damage that may catastrophically disrupt homeostasis
during hibernation.
In North America, bat hibernacula range in temperature from approximately
2–14°C, temperatures all permissive to growth of G. destructans. Within this
temperature range, G. destructans exhibits increasing growth rates with increas-
ing temperature (Figure A4-4), but the fungus does not grow at temperatures
of approximately 20°C or higher. This temperature sensitivity helps to explain
why WNS is observed only among hibernating or recently emerged bats and
why the disease is not diagnosed in bats during their active season when body
temperatures are consistently elevated above those permissive to growth of G.
destructans.
(A)
(B)
Figure A4-3A.eps
bitmap
FIGURE A4-3╇(A) Three little brown bats (Myotis lucifugus) photographed by Alan
Hicks (New York State Department of Environmental Conservation) in Graphite Mine,
New York, in November, 2008. Note the white fungus colonizing the muzzles and nostrils
Figure fungal
of all three bats. Also note the extensive A4-3B.eps
colonization of the skin of the ears and
bitmap
wings of the bat pictured on the right; (B) Periodic acid-Schiff (PAS) stained microscopic
section of wing membrane from a little brown bat with white-nose syndrome collected in
Pennsylvania in February, 2009. Dense colonies of fungal hyphae erode skin and fill the
cup-shaped depressions (arrow). Ulceration of epidermis with penetration and replacement
of subcutaneous tissue (arrow heads) dramatically alters the integrity of wing membrane.
Bar = 25 µm.
APPENDIX A 173
0.8
0.7
Colony Expansion (mm/day)
0.6
0.5
0.4
0.3
0.2
0.1
0
0 5 10 15 20 25
Temperature
FIGURE A4-4╇ Colony expansion rates of Geomyces destructans when grown on corn-
meal agar at 3, 7, 14, and 20°C. The trend line estimates colony expansion rates at tem-
peratures ranging from 3–20°C.
Figure A4-4.eps
redrawn
fungi are more apt to cause fatal diseases in ectothermic, or cold-blooded, organ-
isms such as insects, fish, amphibians, and plants. Bats and other mammals that
hibernate are unique in that they are warm-blooded when metabolically active,
but cold-blooded during hibernation—a period when their metabolism and body
temperatures are dramatically suppressed. Although lowered body temperatures
may predispose torpid bats to infection by G. destructans, the mechanism en-
abling this specific fungus to be a pathogen for bats while other cave-associated
fungi remain innocuous is not known.
How G. destructans kills bats is under active investigation. One possibility
is that fungal infection disrupts how bats behave while hibernating, leading to
more frequent or longer arousals from torpor and thus accelerating usage of fat
reserves. However, fat depletion is not consistently observed among all bats with
WNS. Infected bats also may exhibit other aberrant behaviors midway through
the hibernation season, such as shifting from thermally stable roost sites deep
within hibernacula to areas with more variable temperatures near entrances.
174 FUNGAL DISEASES
Sometimes, they depart early from hibernacula. Thus, exposure to cold could
account for some WNS-associated mortality.
Further, fungal damage to wing membranes, which can account for more
than 85% of the total surface area of a bat, may increase fatality rates. In addition
to the key role that wings play in flight, wing membrane integrity is essential for
maintaining water balance, temperature, blood circulation, and cutaneous respi-
ration. Disrupting any of these functions could increase WNS mortality rates.
As with so many other diseases, the environment affects the progress and
transmission of WNS. Some pathogenic fungi such as Histoplasma capsulatum,
Cryptococcus spp., and Batrachochytrium dendrobatidis can persist in the envi-
ronment without an animal host for survival. This independence contrasts with
host-requiring viruses or other pathogens for which transmission dynamics tend
to moderate as infected hosts are removed from a population. G. destructans
likely does not require bat hosts to survive and can persist in caves by exploiting
other nutrients.
The cool and humid conditions of underground hibernacula provide ideal
environmental conditions for G. destructans or other fungal growth. While most
G. destructans isolates were cultured from skin or fur of bats collected in or near
underground hibernacula during winter, DNA from the same fungus is found
in soil samples from several hibernacula that harbor WNS-infected bats in the
northeastern US. Also, G. destructans has been cultured from soil samples from
hibernacula in three states where WNS occurs, supporting the hypothesis that bat
hibernacula are reservoirs for this pathogen and that bats, humans, or fomites may
transport G. destructans between hibernacula. How temperature and humidity
differences among hibernacula influence G. destructans and WNS is not known.
manage the disease cycle. The sudden and unexpected emergence of WNS exem-
plifies the importance of monitoring, investigating, and responding to emerging
wildlife diseases and the ecological and societal threats that they present.
SUGGESTED READING
Blehert, D. S., A. C. Hicks, M. Behr, C. U. Meteyer, B. M. Berlowski-Zier, E. L. Buckles, J. T.
H. Coleman, S. R. Darling, A. Gargas, R. Niver, J. C. Okoniewski, R. J. Rudd, and W. B.
Stone. 2009. Bat white-nose syndrome: an emerging fungal pathogen? Science 323:227.
Casadevall, A. 2005. Fungal virulence, vertebrate endothermy, and dinosaur extinction: Is there a
connection? Fungal Genet. Biol. 42:98–106.
Cryan, P. M., C. U. Meteyer, D. S. Blehert, and J. G. Boyles. 2010. Wing pathology of white-nose
syndrome in bats suggests life-threatening disruption of physiology. BMC Biol. 8:135.
Desprez-Loustau, M-L., C. Robin, M. Buée, R. Courtecuisse, J. Garbaye, F. Suffert, I. Sache,
and D. M. Rizzo. 2007. The fungal dimension of biological invasions. Trends Ecol. Evol.
22:472–480.
Frick, W. F., J. F. Pollock, A. C. Hicks, K. E. Langwig, D. S. Reynolds, G. G. Turner, C. M.
Butchkoski, and T. H. Kunz. 2010. An emerging disease causes regional population collapse
of a common North American bat species. Science 329:679–682.
Gargas, A., M. T. Trest, M. Christensen, T. J. Volk, and D. S. Blehert. 2009. Geomyces destructans
sp. nov. associated with bat white-nose syndrome. Mycotaxon 108:147–154.
Kunz, T. H. and M. B. Fenton (ed.). 2003. Bat ecology. University of Chicago Press, Chicago.
Lindner, D. L., A. Gargas, J. M. Lorch, M. T. Banik, J. Glaeser, T. H. Kunz, and D. S. Blehert.
2010. DNA-based detection of the fungal pathogen Geomyces destructans in soil from bat
hibernation sites. Mycologia 103:241–246.
Meteyer, C. U., E. L. Buckles, D. S. Blehert, A. C Hicks, D. E. Green, V. Shearn-Bochsler, N.
J. Thomas, A. Gargas, and M. J. Behr. 2009. Pathology criteria for confirming white-nose
syndrome in bats. J. Vet. Diag. Invest. 21:411–414.
Wibbelt, G., A. Kurth, D. Hellmann, M. Weishaar, A. Barlow, M. Veith, J. Prüger, T. Görföl,
T. Grosche, F. Bontadina, U. Zöphel, H.-P. Seidl, P. M. Cryan, and D. S. Blehert. 2010.
White-nose syndrome fungus (Geomyces destructans) in bats, Europe. Emerg. Infect. Dis.
16:1237–1242.
APPENDIX A 177
A5
Abstract
Endothermy and homeothermy are mammalian characteristics whose evolu-
tionary origins are poorly understood. Given that fungal species rapidly lose their
capacity for growth above ambient temperatures, we have proposed that mamma-
lian endothermy enhances fitness by creating exclusionary thermal zones that pro-
tect against fungal disease. According to this view, the relative paucity of invasive
fungal diseases in immunologically intact mammals relative to other infectious
diseases would reflect an inability of most fungal species to establish themselves
in a mammalian host. In this study, that hypothesis was tested by modeling the
fitness increase with temperature versus its metabolic costs. We analyzed the trad-
eoff involved between the costs of the excess metabolic rates required to maintain
a body temperature and the benefit gained by creating a thermal exclusion zone
that protects against environmental microbes such as fungi. The result yields an
optimum at 36.7°C, which closely approximates mammalian body temperatures.
This calculation is consistent with and supportive of the notion that an intrinsic
thermally based resistance against fungal diseases could have contributed to the
success of mammals in the Tertiary relative to that of other vertebrates.
Importance
Mammals are characterized by both maintaining and closely regulating high
body temperatures, processes that are known as endothermy and homeothermy,
18╛╛Originally published as: Bergman, A. and A. Casadevall. 2010. Mammalian Endothermy Op-
timally Restricts Fungi and Metabolic Costs. mBio 1(5): e00212-10. doi:10.1128/mBio.00212-10.
19╛╛Received 17 August 2010 Accepted 11 October 2010 Published 9 November 2010 Citation
Bergman, A., and A. Casadevall. 2010. Mammalian endothermy optimally restricts fungi and meta-
bolic costs. mBio 1(5):e00212-10. doi: 10.1128/mBio.00212-10. Editor Françoise Dromer, Institut
Pasteur Copyright © 2010 Bergman and Casadevall. This is an open-access article distributed under
the terms of the Creative Commons Attribution-Noncommercial-Share Alike 3.0 Unported License,
which permits unrestricted noncommercial use, distribution, and reproduction in any medium, pro-
vided the original author and source are credited. Address correspondence to Arturo Casadevall,
arturo.casadevall@einstein.yu.edu.
20╛╛Department of Systems and Computational Biology, Albert Einstein College of Medicine, Bronx,
respectively. The mammalian lifestyle is energy intensive and costly. The evolu-
tionary mechanisms responsible for the emergence and success of these mamma-
lian characteristics are not understood. This work suggests that high mammalian
temperatures represent optima in the tradeoff between metabolic costs and the
increased fitness that comes with resistance to fungal diseases.
Endothermy and homeothermy are fundamental aspects of mammalian phys-
iology whose evolutionary origin remains poorly understood. Although many ex-
planations have been suggested for the origins of endothermy and homeothermy,
none are fully satisfactory given their high metabolic costs (Kemp, 2008; Ruben,
1995). Furthermore, the factors responsible for the mammalian set point remain
unknown, posing the additional question of why mammals are so hot. Recently,
the observation that fungal diseases are common in plants and insects but rare in
mammals, combined with the thermal susceptibility of fungi, led to the proposal
that mammalian endothermy and homeothermy create a thermal exclusionary
zone that protects mammals against mycoses (Robert and Casadevall, 2009).
Endothermy was also suggested to have provided a fitness advantage in the
fungal bloom that followed the end of the Cretaceous such that it could have
contributed to the success of mammals in the Tertiary (Casadevall, 2005; Robert
and Casadevall, 2009).
Assuming that a relationship exists between endothermy and reduced sus-
ceptibility to certain classes of microbes, we hypothesized a tradeoff relationship
whereby the high costs of endothermy were mitigated by protection against in-
fectious diseases. In other words, we posited that increases in body temperature
would protect against microbes by creating a thermal exclusionary zone but that
such increases would be increasingly costly with regard to metabolic rates as the
host body temperature diverged from ambient temperatures. Given that there is
robust information on fungal thermal tolerances (Robert and Casadevall, 2009),
we decided to test this hypothesis by attempting to identify body temperatures
that confer maximal fitness for certain metabolic rates.
To address this question, we propose a first-order model wherein a tradeoff
exists between the excess metabolic rates required to maintain a body tempera-
ture, T, and the benefit gained by protection against deleterious microbes because
of the creation of a thermal exclusion zone. Metabolism, the exchange of energy
between the organism and its environment, as well as the transformation of that
energy to material within an organism, is affected by two main factors, body
mass, M, and body temperature, T. Due to the fractal nature of transport networks,
that is, vessel architecture and branching (Gillooly et al., 2001; Savage et al.,
2008), over ontogeny, the resting metabolic rate, Brest, scales with body mass, m,
as Brest = B0m3⁄4, where B0 is a normalization constant for a given taxon. Also, the
normalization coefficient, B0, exponentially increases with body temperature B0
~ e–E0/KT, where E0 is the average activation energy for the rate-limiting enzyme-
catalyzed biochemical reactions of metabolism (ca. 0.65 eV), K is Boltzmann’s
constant (8.62 × 10–5eV/K), and T is body temperature (Brown et al., 2004;
APPENDIX A 179
Gillooly et al., 2001). The scaling relationship between resting metabolic rate
and body mass, ∝m3⁄4, has been predicted from allometric theories and supported
by data on a diverse set of organisms, including mammals, birds, fish, and mol-
lusks (Brody, 1964; Moses et al., 2008; Savage et al., 2004; West et al., 1997).
As can be seen from the formulas above, body temperature affects the metabolic
rate through its effects on rates of biochemical reaction kinetics according to
Boltzmann’s factor, e–Ei⁄kT, where T is measured in kelvins (absolute temperature).
The resting metabolic rate, Brest, is proportional to the product of these two ef-
fects and again has been shown to be well approximated, within a biologically
relevant temperature range (0°C to 40°C), as B(T) ∝e–Ei⁄kTm3/4 (Gillooly, 2001).
The first part of our analysis examined the excess cost for an organism of body
mass m to maintain a body temperature T (assuming no dependence of body mass
on temperature).
In the second part of our analysis, the benefit, noted here as F(T), is calcu-
lated as the reduction in the number of fungal species capable of infecting a host;
this number is reduced approximately by s ≈ 6% for every degree Celsius in the
temperature range of 27°C to 40°C (Robert and Casadevall, 2009). The increased
benefit of the successive elimination of fungal species can thus be expressed as
F(T) ∝F0[1 – (1 – s)T], where F0 is a constant scaling factor. The quantity W(T)
= F(T)⁄B(T) can represent the balance between cost and benefit; thus, W(T) can be
viewed as the total fitness of an organism as a function of its body temperature.
Within the biologically relevant temperature range, the proposed fitness measure
reaches its maximum at approximately 37°C (Fig. A5-1). Note that in this for-
mulation, the optimal body temperature, where W(T) attains its maximum value,
does not depend on the organism’s body mass. Furthermore, the one parameter
that is determined from biological observation is the reduction in the number of
fungal species capable of infecting a host; thus, to determine our model’s de-
pendence on this parameter, we calculated the optimal temperature over a wide
range of possible reduction percentages, i.e., 4% to 8%. In this range, the optimal
temperature was found to remain in a tight range of less than 2°C, from 37.7°C to
35.9°C, respectively, which is still within the biologically relevant range of mam-
malian body temperatures. The insensitivity of the model to its only parameter
further strengthens our hypothesis.
In summary, we present a minimal, parsimonious model to account for the
cost of maintaining a high body temperature in mammalian organisms. A body
temperature of 36.7°C maximizes fitness by restricting the growth of most fun-
gal species relative to its metabolic cost. Our model suggests that no additional
elaborations are required to explain the evolution of endothermy other than the
tradeoff between protection against environmentally acquired microbial dis-
eases and the cost of metabolism. Although we cannot rule out the possibility
that this body temperature optimum arose by some remarkable coincidence, we
think this highly unlikely because it emerges from considering two unrelated
processes, fungal thermal tolerance and mammalian metabolic costs. Nonethe-
180 FUNGAL DISEASES
less, we acknowledge that similar temperature optima might emerge from other
considerations. For example, the specific heat capacity of water has a minimum
at 36°C, and if the efficiency of metabolic processes is related to heat capacity,
then using this parameter as the optimality criterion may result in a similar range
of solutions. Nevertheless, we note the internal consistency in the theme that
fungal diseases are rare in immunologically intact mammals and the tradeoff
between increased fitness and metabolic costs closely approximates mammalian
body temperatures.
Acknowledgements
Aviv Bergman is supported by 5P01AG027734-04 and 5R01AG028872-04.
Arturo Casadevall is supported by AI33774-11, HL59842-07, AI33142-11,
AI52733-02, and U54-AI057158-Lipkin.
References
Brody, S. 1964. Bioenergetics and growth. Hafner, New York, NY.
Brown, J. H., J. F. Gillooly, A. P. Allen, V. M. Savage, and G. B. West. 2004. Toward a metabolic
theory of ecology. Ecology 85:1771–1789.
APPENDIX A 181
Casadevall, A. 2005. Fungal virulence, vertebrate endothermy, and dinosaur extinction: is there a
connection? Fungal Genet. Biol. 42:98–106.
Gillooly, J. F., J. H. Brown, G. B. West, V. M. Savage, and E. L. Charnov. 2001. Effects of size and
temperature on metabolic rate. Science 293:2248–2251.
Kemp, T. S. 2008. The origin of mammalian endothermy: a paradigm for the evolution of complex
biological structure. Zool. J. Linn. Soc. 147:473–488.
Moses, M. E., C. Hou, W. H. Woodruff, G. B. West, J. C. Nekola, W. Zuo, and J. H. Brown. 2008.
Revisiting a model of ontogenetic growth: estimating model parameters from theory and data.
Am. Nat. 171:632–645.
Robert, V. A., and A. Casadevall. 2009. Vertebrate endothermy restricts most fungi as potential patho-
gens. J. Infect. Dis. 200:1623–1626.
Ruben, J. 1995. The evolution of endothermy in mammals and birds: from physiology to fossils.
Annu. Rev. Physiol. 57:69–95.
Savage, V. M., E. J. Deeds, and W. Fontana. 2008. Sizing up allometric scaling theory. PLoS Comput.
Biol. 4:e1000171.
Savage, V. M., J. F. Gillooly, W. H. Woodruff, G. B. West, A. P. Allen, B. J. Enquist, and J. H. Brown.
2004. The predominance of quarter-power scaling in biology. Funct. Ecol. 18:257–282.
West, G. B., J. H. Brown, and B. J. Enquist. 1997. A general model for the origin of allometric scaling
laws in biology. Science 276:122–126.
A6
22╛╛Vincent A.Robert and Arturo Casadevall, “A Vertebrate Endothermy Restricts Most Fungi as Po-
tential Pathogens”, Journal of Infectious Diseases, 2009, Vol. 200, Iss. 10, pp. 1623–1626. Reprinted
by permission of Oxford University Press.
23╛╛Received 23 May 2009; accepted 18 June 2009; electronically published 14 October 2009. Po-
tential conflicts of interest: none reported. Financial support: National Institutes of Health (awards
5R01AI033774, 5R01HL059842, and 2U54AI057158). Reprints or Correspondence: Dr Arturo Casa-
devall, Department of Medicine, Albert Einstein College of Medicine, Yeshiva University, 1300
Morris Park Ave, Bronx, NY 10461 (arturo.casadevall@einstein.yu.edu).
╛╛The Journal of Infectious Diseases 2009;200:000–000
╛╛© 2009 by the Infectious Diseases Society of America. All rights reserved. 0022-1899/2009/
20010-00XX$15.00
╛╛DOI: 10.1086/644642
24╛╛Centraalbureau voor Schimmelcultures, Utrecht, the Netherlands.
25╛╛Department of Microbiology and Immunology and Division of Infectious Diseases, Department
of Medicine, Albert Einstein College of Medicine, Yeshiva University, Bronx, New York.
182 FUNGAL DISEASES
Methods
A total of 4802 fungal strains belonging to 144 genera in the Centraalbureau
voor Schimmelcultures (Utrecht) collection were tested for growth at 4°C, 12°C,
15°C, 18°C, 21°C, 25°C, 30°C, 35°C, 37°C, 40°C, 42°C, and 45°C. Strains were
grown for times ranging from a few days to a few weeks on the most suitable
medium, generally glucose–peptone–yeast extract agar, potato-dextrose agar, or
yeast extract–malt extract agar. Growth was considered positive when a colony
was visible without magnification. The strain set included Ascomycetes and
Basidiomycetes but excluded Zygomycetes, which is not in the yeast database.
The culture deposit records were reviewed to identify the isolation source.
Fungi isolated from flowers, grains, and herbal exudates were grouped under
plant isolates. Animal isolates were classified depending on whether they origi-
nated from endothermic (mammals and birds) or ectothermic (insects, nematodes,
fishes, and crustaceans) species. Another group comprised isolates from nonliving
environmental sources, which included predominantly soils; this group is referred
to as soil isolates. These groups were compared for thermal tolerance at 2 tem-
APPENDIX A 183
peratures, 25°C and 37°C, which reflect ambient and mammalian temperatures,
respectively.
To test the significance of the difference in growth patterns between fungal
strains isolated from different groups, we calculated the test statistics
) )
z = ( p 1 − p2 / P × (1 − P ) × (1 / n1 + 1 / n 2 ,
where p1 and p2 are the observed sample proportions for each group of fungal
strains at a given temperature, n1 and n2 are the size of the 2 groups under com-
parison, and
FIGURE A6-1╇ Frequency histogram of thermal growth tolerance for 4802 fungal strains
(bars). Lines connect percentagesFigure
for 49 mammalian
A6-1.eps (blue) and 12 bird (red) species core
temperatures. Obtained from McNab (1970).
bitmap
TABLE A6-1╇ Growth Tolerances for Fungi from Soils, Animals, and Plants at
2 Temperatures
Isolate Growth P valuesb
Origin, host type Yes No Unknowna Total P1 P2 P3
at 25°C
Soils, NA 657 42 7 706
Plant, ectotherm 1108 30 5 1143 <.001
Animal
Ectotherm 490 0 6 496 <.001 .29
Endotherm 661 5 9 675 <.001 .214 .263
at 37°C
Soils, NA 146 535 15 706
Plant, ectotherm 304 871 22 1143 .292
Animal
Ectotherm 193 284 19 496 <.001 .004
Endotherm 466 202 7 605 <.001 <.001 <.001
NOTE. NA, not applicable
a Refers to a small no. of isolates for which the temperature growth data was not complete.
b P1 refers to the comparison of isolates from soils, P2 refers to the comparison versus plant iso-
lates, and P3 refers to the comparison between isolates from ectothermic and endothermic animals.
APPENDIX A 185
fungi that grew was much higher for isolates from endothermic animals than
from ectothermic animals. The proportions of Ascomycetes and Basidiomycetes
fungi in each group were comparable, except for ectothermic hosts, which yielded
predominantly Ascomycetes fungi.
Isolates from ectothermic hosts (such as plants and insects) were significantly
more thermotolerant than isolates from soils. A significantly greater percentage of
fungal strains from insects grew at 37°C relative to those recovered from plants,
possibly reflecting the fact that insects can increase their temperature through
behavioral fevers that increase survival after fungal infection (Thomas and Blan-
ford, 2003). However, this explanation is unlikely to apply to plants, which have
much lower metabolic rates. Since thermal tolerance must be associated with nu-
merous metabolic changes that mitigate fungal damage, the association between
greater thermotolerance and plant pathogenicity could mirror adaptation to sur-
vival in a host with potent antifungal defenses, raising the tantalizing possibility
that selection pressures by virulence may contribute to thermal stability and vice
versa. In this regard, we note that Hsp90 orchestrates morphogenesis in Candida
albicans (Shapiro et al., 2009), thus providing a molecular association for heat
shock and a virulence-related phenotype that may be conserved in other fungi.
A survey of the fungal genera represented in our sample collection revealed
differences in the percentage of isolates capable of growth at 37°C. All genera
studied included some thermotolerant species, as defined by their ability to grow
at 37°C, but there were large differences in the percentage of species within
each genera. Thermotolerant genera included those from both Ascomycetes and
Basidiomycetes, but basidiomycetous genera were disproportionately more com-
mon among the thermotolerant genera (P < .001, Fisher exact test). The strains
grouped within the sexually related basidiomycetous genera Filobasidiella (a
telemorph of Cryptococcus) and Cryptococcus (an anamorph of Filobasidiella)
included comparable numbers of thermotolerant species (61% among 116 strains
and 53% among 287 strains, respectively). These data suggest an association
between phylogeny and thermotolerance.
The capacity for thermotolerance was interspersed among Ascomycetes and
Basidiomycetes, suggesting that it may have emerged independently several
times in evolution. Alternatively, thermotolerance may be an ancient fungal trait
that was lost by those species that cannot grow at 37°C. In this regard, we note
that the climate for much of Earth’s history was much warmer than in recent
geologic epochs, having cooled by ~5°C during the Eocene-Oligocene transition
~34 million years ago (Liu et al., 2009). The fact that thermotolerance is a com-
plex trait that can be lost by a single mutation, as demonstrated by laboratory-
generated temperature-sensitive mutants, makes the explanation of a retained
phenotype attractive.
Our results may be relevant to the ongoing debate on the origin and func-
tion of endothermy, homeothermy, and fever, each a major unsolved problem in
vertebrate physiology (Kemp, 2008; Ruben, 1995). There is no consensus as to
186 FUNGAL DISEASES
References
Blehert DS, Hicks AC, Behr M, et al. Bat white-nose syndrome: an emerging fungal pathogen? Sci-
ence 2009; 323:227.
Casadevall A. Fungal virulence, vertebrate endothermy, and dinosaur extinction: is there a connec-
tion? Fungal Genet Biol 2005; 42:98–106.
Casadevall A, Pirofski LA. Accidental virulence, cryptic pathogenesis, martians, lost hosts, and the
pathogenicity of environmental microbes. Eukaryot Cell 2007; 6:2169–74.
Dromer F, Mathoulin S, Dupont B, Letenneur L, Ronin O. Individual and environmental factors
associated with infection due to Cryptococcus neoformans serotype D. Clin Infect Dis 1996;
23:91–6.
Hawksworth DL, Rossman AY. Where are all the undescribed fungi? Phytopathology 1997; 87:888–91.
Kemp TS. The origin of mammalian endothermy: a paradigm for the evolution of complex biological
structure. Zool J Linn Soc 2008; 147:473–88.
Kwon-Chung KJ, Bennett JE. Medical mycology. Philadelphia: Lea & Febiger, 1992.
Liu Z, Pagani M, Zinniker D, et al. Global cooling during the eocene-oligocene climate transition.
Science 2009; 323:1187–90.
Martinez LR, Garcia-Rivera J, Casadevall A. Cryptococcus neoformans var. neoformans (serotype
D) strains are more susceptible to heat than C. neoformans var. grubii (serotype A) strains. J
Clin Microbiol 2001; 39:3365–7.
McNab BK. Body weight and the energetics of temperature regulation. J Exp Biol 1970; 53:329–48.
Obendorf DL, Peel BF, Munday BL. Mucor amphibiorum infection in platypus (Ornithorhynchus
anatinus) from Tasmania. J Wildl Dis 1993; 29:485–7.
Perfect JR, Lang SDR, Durack DT. Chronic cryptococcal meningitis. Am J Path 1980; 101:177–93.
188 FUNGAL DISEASES
Ruben J. The evolution of endothermy in mammals and birds: from physiology to fossils. Annu Rev
Physiol 1995; 57:69–95.
Shapiro RS, Uppuluri P, Zaas AK, et al. Hsp90 orchestrates temperature-dependent Candida albicans
morphogenesis via Ras1-PKA signaling. Curr Biol 2009; 19:621–9.
Thomas MB, Blanford S. Thermal biology in insect-pathogen interactions. Trends Ecol Evol 2003;
18:344–50.
A7
Number/
Percentage
173/48.3%
21/5.9%
13/3.6%
1/0.3%
42/11.7%
21/5.9%
87/24.3%
the United States, the U.S. Fish and Wildlife Service is the agency responsible for
protecting wildlife, but if the threat is an emerging pathogen, this agency has little
capacity for outbreak investigation and control. Similarly, the national agency
responsible for funding ecological research in the United States is the National
Science Foundation (NSF). The National Institutes of Health (NIH) oversees a
broad range of issues, from infections to organ dysfunction to mental health and
other areas. The U.S. Department of Agriculture oversees agricultural health. At
the international scale, the intergovernmental agency to protect wildlife is the
International Union for the Conservation of Nature (IUCN), whereas the global
health agenda falls under the World Health Organization (WHO), agricultural
health under the Food and Agriculture Organization (FAO), and trade-related
disease issues under the World Organisation for Animal Health (OIE). This siloed
approach is followed by many countries globally; they tend to have separate
ministries for health, agriculture, trade, and environment/forestry/wildlife. These
approaches work well until the threats to human health cross these jurisdictional
boundaries. With emerging diseases, they have done so repeatedly. For example,
the emergence of severe acute respiratory syndrome involved wildlife reservoir
APPENDIX A 191
species (Li et al., 2005), the national and international trade in hunted and farmed
wildlife and livestock (Xu et al., 2004), and international travel and migration
(Anderson et al., 2004). Likewise, the global emergence of amphibian chytrid-
iomycosis has been linked to trade and climate change (Lips et al., 2008), and
involves the medical industry (Weldon et al., 2004), the production of amphibians
for food (Schloegel et al., 2009), and introduced or invasive species (Kilpatrick
et al., 2010).
One simple approach to overcoming these challenges is to encourage cross-
disciplinary, cross-agency collaboration. This approach to research and policy
has been led by the fields of “conservation medicine” (Daszak et al., 2004), One
Health (Karesh and Cook, 2005), and EcoHealth (Daszak, 2009; Wilcox and
Daszak, 2006). In the United States, some efforts have successfully bridged the
funding gap between NIH and NSF, notably the Ecology of Infectious Diseases
program launched jointly by the NSF and the NIH John E. Fogarty International
Center in 2000 (Scheiner and Rosenthal, 2006). Likewise, there is a unique
U.S. federal agency with a specific remit to address with wildlife diseases, the
National Wildlife Health Center (NWHC) (Fleischli et al., 2004; Skerratt et al.,
2005). The NWHC has been conducting surveillance, monitoring, investigation,
research, and response on wildlife diseases for 35 years, and is registered with
the Centers for Disease Control and Prevention Select Agent Program, marking
it as a laboratory of significant relevance to human and livestock as well as wild-
life health. It has a sophisticated network of laboratories, including Biosecurity
Level 3 biocontainment labs, necropsy suites, and isolation rooms. In addition,
it publishes quarterly reports of mortality investigations, and acts as a national
focal point for similar activities in universities and NGOs. At the intergovern-
mental scale, there has been a recent flurry of activity to bring together agencies
around the One Health agenda, including formal links among the OIE, FAO, and
WHO, which originated from their collaborative efforts to tackle avian influenza
(Anderson et al., 2010). Additionally, wildlife health has two significant nuclei
within the United Nations system.
First, in the IUCN, there is the Species Survival Commission Wildlife Health
Specialist Group (http://www.iucn.org/about/work/programmes/species/about_
ssc/specialist_groups/specialist_group_pprofiles/veterinary_sg_profile/), which
has a network of more than 400 wildlife veterinarians and researchers globally.
Second, the OIE has a Working Group on Wildlife Diseases (http://web.oie.int/
wildlife/eng/en_wildlife.htm), which has operated for more than 15 years and
advises the OIE on wildlife health issues.
These initiatives have begun to bring diverse disciplines together to under-
stand the drivers and impacts of wildlife EIDs, and to conduct effective surveil-
lance and control of wildlife as reservoirs for human EIDs. However, they could
be improved significantly with some simple approaches. First, within the United
States, the government has the capacity to form interagency task forces for spe-
cific issues that cross agency mandates. In a previous administration, the complex
issue of amphibian declines was addressed with the formation of the Interagency
192 FUNGAL DISEASES
Taskforce on Amphibian Declines and Deformities. Similar task forces are likely
to be useful to address the need for better surveillance of the wildlife trade for
pathogens (Smith et al., 2009a), or the threat of white-nose syndrome in bats.
Second, at a global scale, strengthening of laboratory capacity and personnel for
wildlife diseases, and support for One Health approaches from the development
community, could be extremely useful in fostering linkages among disparate
ministries, universities, NGOs, and others. Recently, the U.S. Agency for Inter-
national Development launched an Emerging Pandemic Threats program that
specifically adopted a One Health approach to build capacity in the regions where
emerging zoonoses most commonly originate (http://www.usaid.gov/our_work/
global_health/home/News/ai_docs/ept_brochure.pdf) (Daszak, 2009). This in-
cludes specific collaboration among human and veterinary medical scientists;
ministries of health, agriculture, and environment; and OIE, FAO, and WHO.
References
Aguirre, A. A., and G. M. Tabor. 2008. Global factors driving emerging infectious diseases: Impact
on wildlife populations. Annals of the New York Academy of Sciences 1149:1–3.
Anderson, P. K., A. A. Cunnigham, N. G. Patel, F. J. Morales, P. R. Epstein, and P. Daszak. 2004.
Emerging infectious diseases of plants: Crop homogeneity, pathogen pollution and climate
change drivers. Trends in Ecology and Evolution 19(10):535–544.
Anderson, R. M., C. Fraser, A. C. Ghani, C. A. Donnelly, S. Riley, N. M. Ferguson, G. M. Leung,
T. H. Lam, and A. J. Hedley. 2004. Epidemiology, transmission dynamics and control of SARS:
The 2002–2003 epidemic. Philosophical Transactions of the Royal Society of London Series B:
Biological Sciences 359:1091–1105.
Anderson, T., et al. 2010. FAO–OIE–WHO Joint Technical Consultation on avian influenza at the
human–animal interface. Influenza and Other Respiratory Viruses 4:1–29.
Berger, L., R. Speare, P. Daszak, D. E. Green, A. A. Cunningham, C. L. Goggin, R. Slocombe, M. A.
Ragan, A. D. Hyatt, K. R. McDonald, H. B. Hines, K. R. Lips, G. Marantelli, and H. Parkes.
1998. Chytridiomycosis causes amphibian mortality associated with population declines in the
rain forests of Australia and Central America. Proceedings of the National Academy of Sciences,
USA 95:9031–9036.
27╛╛Markets which sell live wildlife (often mixed with livestock) are called “Wetmarkets,” particu-
Blaustein, A. R. 1994. Chicken little or Nero’s fiddle? A perspective on declining amphibian popula-
tions. Herpetologica 50:85–97.
Blehert, D. S., A. C. Hicks, M. Behr, C. U. Meteyer, B. M. Berlowski-Zier, E. L. Buckles, J. T.
Coleman, S. R. Darling, A. Gargas, R. Niver, J. C. Okoniewski, R. J. Rudd, and W. B. Stone.
2009. Bat white-nose syndrome: An emerging fungal pathogen? Science 323:227.
Carlton, J. T. 1993. Neoextinctions of marine-invertebrates. American Zoologist 33:499–509.
Carlton, J. T., G. J. Vermeij, D. R. Lindberg, D. A. Carlton, and E. C. Dudley. 1991. The 1st historical
extinction of a marine invertebrate in an ocean-basin—the demise of the eelgrass limpet Lottia-
alveus. Biological Bulletin 180:72–80.
Chaturvedi, V., and S. Chaturvedi. 2011. Editorial: What is in a name? A proposal to use geomycosis
instead of white nose syndrome (WNS) to describe bat infection caused by Geomyces destruc-
tans. Mycopathologia 171:231–233.
Crawford, A. J., K. R. Lips, and E. Bermingham. 2010. Epidemic disease decimates amphibian
abundance, species diversity, and evolutionary history in the highlands of central Panama.
Proceedings of the National Academy of Sciences, USA 107:13777–13782.
Cunningham, A. A. 2005. A walk on the wild side—emerging wildlife diseases: They increasingly
threaten human and animal health. British Medical Journal 331:1214–1215.
Daszak, P. 2009. A call for “smart surveillance”: A lesson learned from H1N1. Ecohealth 6:1–2.
Daszak, P., and A. A. Cunningham. 1999. Extinction by infection. Trends in Ecology & Evolution
14:279.
Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2000. Emerging infectious diseases of wildlife—
threats to biodiversity and human health. Science 287:443–449.
Daszak, P., G. M. Tabor, A. M. Kilpatrick, J. Epstein, and R. Plowright. 2004. Conservation medi-
cine and a new agenda for emerging diseases. Annals of the New York Academy of Sciences
1026:1–11.
Daszak, P., D. E. Scott, A. M. Kilpatrick, C. Faggioni, J. W. Gibbons, and D. Porter. 2005. Amphibian
population declines at Savannah River site are linked to climate, not chytridiomycosis. Ecology
86:3232–3237.
Deem, S. L., W. B. Karesh, and W. Weisman. 2001. Putting theory into practice: Wildlife health in
conservation. Conservation Biology 15:1224–1233.
Fischer, J. R., D. E. Stallknecht, M. P. Luttrell, A. A. Dhondt, and K. A. Converse. 1997. Mycoplas-
mal conjunctivitis in wild songbirds: The spread of a new contagious disease in a mobile host
population. Emerging Infectious Diseases 3:69–72.
Fleischli, M. A., J. C. Franson, N. J. Thomas, D. L. Finley, and W. Riley. 2004. Avian mortality
events in the United States caused by anticholinesterase pesticides: A retrospective summary of
National Wildlife Health Center records from 1980 to 2000. Archives of Environmental Con-
tamination and Toxicology 46:542–550.
Frick, W. F., J. F. Pollock, A. C. Hicks, K. E. Langwig, D. S. Reynolds, G. G. Turner, C. M.
Butchkoski, and T. H. Kunz. 2010. An emerging disease causes regional population collapse of
a common North American bat species. Science 329(5992):679–682.
Jones, K. E., N. G. Patel, M. A. Levy, A. Storeygard, D. Balk, J. L. Gittleman, and P. Daszak. 2008.
Global trends in emerging infectious diseases. Nature 451:990–994.
Karesh, W. B., and R. A. Cook. 2005. The human–animal link, one world–one health. Foreign Af-
fairs 84:38–50.
Karesh, W. B., R. A. Cook, E. L. Bennett, and J. Newcomb. 2005. Wildlife trade and global disease
emergence. Emerging Infectious Diseases 11:1000–1002.
Kilpatrick, A. M., C. J. Briggs, and P. Daszak. 2010. The ecology and impact of chytridiomycosis: An
emerging disease of amphibians. Trends in Ecology & Evolution 25:109–118.
King, D. A., C. Peckham, J. K. Waage, J. Brownlie, and M. E. J. Woolhouse. 2006. Infectious dis-
eases: Preparing for the future. Science 313:1392–1393.
APPENDIX A 195
Li, W. D., Z. Shi, M. Yu, W. Ren, C. Smith, J. H. Epstein, H. Wang, G. Crameri, Z. Hu, H. Zhang, J.
Zhang, J. McEachern, H. Field, P. Daszak, B. T. Eaton, S. Zhang, and L. F. Wang. 2005. Bats
are natural reservoirs of SARS-like coronaviruses. Science 310:676–679.
Lips, K. R., F. Brem, R. Brenes, J. D. Reeve, R. A. Alford, J. Voyles, C. Carey, L. Livo, A. P. Pessier,
and J. P. Collins. 2006. Emerging infectious disease and the loss of biodiversity in a neotropical
amphibian community. Proceedings of the National Academy of Sciences, USA 103:3165–3170.
Lips, K. R., J. Diffendorfer, J. R. Mendelson, and M. W. Sears. 2008. Riding the wave: Reconciling
the roles of disease and climate change in amphibian declines. PLoS Biology 6:441–454.
Mahy, B. W. J., and C. C. Brown. 2000. Emerging zoonoses: Crossing the species barrier. J Rev Sci
Tech OIE 19:33–40.
Mendelson, J. R., K. R. Lips, R. W. Gagliardo, G. B. Rabb, J. P. Collins, J. E. Diffendorfer, P.
Daszak, D. R. Ibáñez, K. C. Zippel, D. P. Lawson, K. M. Wright, S. N. Stuart, C. Gascon, H. R.
da Silva, P. A. Burrowes, R. L. Joglar, E. La Marca, S. Lötters, L. H. du Preez, C. Weldon, A.
Hyatt, J. V. Rodriguez-Mahecha, S. Hunt, H. Robertson, B. Lock, C. J. Raxworthy, D. R. Frost,
R. C. Lacy, R. A. Alford, J. A. Campbell, G. Parra-Olea, F. Bolaños, J. J. Domingo, T. Halliday,
J. B. Murphy, M. H. Wake, L. A. Coloma, S. L. Kuzmin, M. S. Price, K. M. Howell, M. Lau,
R. Pethiyagoda, M. Boone, M. J. Lannoo, A. R. Blaustein, A. Dobson, R. A. Griffiths, M. L.
Crump, D. B. Wake, and E. D. Brodie, Jr. 2006. Biodiversity: Confronting amphibian declines
and extinctions. Science 313:48.
Miller, M. W., and E. S. Williams. 2004. Chronic wasting disease of cervids. Mad Cow Disease and
Related Spongiform Encephalopathies 284:193–214.
Nettles, V. F. 1996. Reemerging and emerging infectious diseases: Economic and other impacts on
wildlife—Transport of animals sometimes spreads infections, while other outbreaks are a mys-
tery. ASM News 62:589–591.
Neumann, G., T. Noda, and Y. Kawaoka. 2009. Emergence and pandemic potential of swine-origin
H1N1 influenza virus. Nature 459:931–939.
Pavlin, B. I., L. M. Schloegel, and P. Daszak. 2009. Risk of importing zoonotic diseases through
wildlife trade, United States. Emerging Infectious Diseases 15:1721–1726.
Pechmann, J. H. K., D. E. Scott, R. D. Semlitsch, J. P. Caldwell, L. J. Vitt, and J. W. Gibbons. 1991.
Declining amphibian populations: The problem of separating human impacts from natural
fluctuations. Science 253:892–895.
Robinson, R. A., B. Lawson, M. P. Toms, K. M. Peck, J. K. Kirkwood, J. Chantrey, I. R. Clatworthy,
A. D. Evans, L. A. Hughes, O. C. Hutchinson, S. K. John, T. W. Pennycott, M. W. Perkins, P. S.
Rowley, V. R. Simpson, K. M. Tyler, and A. A. Cunningham. 2010. Emerging infectious disease
leads to rapid population declines of common British birds. PLoS ONE 5(8):1–12.
Scheiner, S. M., and J. P. Rosenthal. 2006. Ecology of infectious disease: Forging an alliance. Eco-
health 3:204–208.
Schloegel, L. M., J. M. Hero, L. Berger, R. Speare, K. McDonald, and P. Daszak. 2006. The decline
of the sharp-snouted day frog (Taudactylus acutirostris): The first documented case of extinction
by infection in a free-ranging wildlife species? Ecohealth 3:35–40.
Schloegel, L. M., A. M. Picco, A. M. Kilpatrick, A. J. Davies, A. G. Hyatt, and P. Daszak. 2009.
Magnitude of the U.S. trade in amphibians and presence of Batrachochytrium dendrobatidis
and ranavirus infection in imported North American bullfrogs (Rana catesbeiana). Biological
Conservation 142:1420–1426.
Schloegel, L. M., P. Daszak, A. A. Cunningham, R. Speare, and B. Hill. 2010. Two amphibian dis-
eases, chytridiomycosis and ranaviral disease, are now globally notifiable to the World Organi-
zation for Animal Health (OIE): An assessment. Diseases of Aquatic Organisms 92:101–108.
Schwartz, M. W., S. M. Hermann, and C. S. Vogel. 1995. The catastrophic loss of Torreya-taxifolia—
assessing environmental induction of disease hypotheses. Ecological Applications 5:501–516.
Schwartz, M. W., S. M. Hermann, and P. J. van Mantgem. 2000. Estimating the magnitude of decline
of the Florida torreya (Torreya taxifolia Arn.). Biological Conservation 95:77–84.
196 FUNGAL DISEASES
Sigurdson, C. J. 2008. A prion disease of cervids: Chronic wasting disease. Veterinary Research
39–41.
Skerratt, L. F., J. C. Franson, C. U. Meteyer, and T. E. Hollmen. 2005. Causes of mortality in sea ducks
(Mergini) necropsied at the USGS-National Wildlife Health Center. Waterbirds 28:193–207.
Skerratt, L. F., L. Berger, R. Speare, S. Cashins, K. Raymond McDonald, A. D. Phillott, H. B. Hines,
and N. Kenyon. 2007. Spread of chytridiomycosis has caused the rapid global decline and
extinction of frogs. Ecohealth 4:125–134.
Smith, G. J. D., D. Vijaykrishna, J Bahl, S. J. Lycett, M. Worobey, O. G. Pybus, S. K. Ma, C. L.
Cheung, J. Raghwani, S. Bhatt, J. S. Peiris, Y. Guan, and A. Rambaut. 2009b. Origins and evolu-
tionary genomics of the 2009 swine-origin H1N1 influenza A epidemic. Nature 459:1122–1126.
Smith, K. F., M. Behrens, L. M. Schloegel, N. Maranao, S. Burgiel, and P. Daszak. 2009a. Reducing
the risks of the wildlife trade. Science 324:594–595.
Taylor, L. H., S. M. Latham, and M. E. Woolhouse. 2001. Risk factors for human disease emergence.
Philosophical Transactions of the Royal Society of London B: Biological Sciences 356:983–989.
Weldon, C., L. H. du Preez, A. D. Hyatt, R. Muller, and R. Speare. 2004. Origin of the amphibian
chytrid fungus. Emerging Infectious Diseases 10:2100–2105.
WHO gets mixed reviews for H1N1 response. 2011. Science 331:1371–1371.
Wilcox, B. A., and P. Daszak. 2006. Launching the International EcoHealth Association. EcoHealth
3:125–126.
Wildlife and emerging disease. 2009. Veterinary Record 165:458–459.
Williams, E. S., T. Yuill, M. Artois, J. Fischer, and S. A. Haigh. 2002. Emerging infectious diseases in
wildlife. Revue Scientifique Et Technique De L Office International Des Epizooties 21:139–157.
Woolhouse, M. E. J., and S. Gowtage-Sequeria. 2005. Host range and emerging and re-emerging
pathogens. Emerging Infectious Diseases 11:1842–1847.
Xu, R. H., J. F. He, M. R. Evans, G. W. Peng, H. E. Field, D. W. Yu, C. K. Lee, H. M. Luo, W. S.
Lin, P. Lin, L. H. Li, W. J. Liang, J. Y. Lin, and A. Schnur. 2004. Epidemiologic clues to SARS
origin in China. Emerging Infectious Diseases 10:1030–1037.
A8
Introduction
Human disease resulting from infection by Coccidioides spp. was first rec-
ognized late in the 19th century. Since then, with more information and chang-
ing demographics, our understanding of this problem and our perception of its
importance has evolved in many ways (Galgiani, 2007). First thought of as a rare
28╛╛Valley
Fever Center for Excellence, University of Arizona College of Medicine.
29╛╛Correspondence
and current address: John N. Galgiani, M.D.; Professor, University of Arizona
College of Medicine; Director, Valley Fever Center for Excellence; P.O. Box 245215, Tucson, AZ
85724; Tel.: 520-626-4968; Fax: 520-626-4971; e-mail: spherule@u.arizona.edu.
APPENDIX A 197
methods were available to do this, methods for treating the soil exist that would
minimize or prevent this exposure from happening. Advances in this area would
have practical public health benefits.
patients lost more than a month of work; a quarter of patients needed 10 or more
physician visits; and 40 percent of patients required hospitalization for their
illness (Tsang et al., 2010). The same report shows Arizona hospital costs for
coccidioidomycosis were more than $86 million. Extrapolating from these costs,
estimates that include all outpatient medical care, often lasting for years if not
entire lives, could easily reach a quarter of a billion dollars.
As significant as these findings are, other projections suggest that the actual
number of persons seeking medical attention for coccidioidal infection is several
times greater than those diagnosed and included in state public health statistics.
One study found that only 3–13 percent of patients with pneumonia in Phoenix
were tested for coccidioidomycosis (Chang et al., 2008). By contrast, a prospec-
tive study in Tucson in which patients with community-acquired pneumonia were
tested for coccidioidomycosis demonstrated that nearly a third of these subjects
had a coccidioidal infection (Valdivia et al., 2006).
State statistics show case rates for college-age persons in Pima County to be
from 34 to 48 cases per 100,000 annually. However, recent surveillance of schol-
arship athletes at the University of Arizona, which is in Pima County, indicated
374 cases per 100,000 (Stern and Galgiani, 2010). Further analysis suggested that
the most important reason for this much higher case rate was that the athletes
received many more serologic tests for coccidioidomycosis. Evidently, more pa-
tients would be accurately identified as to the true cause of their illness if patients
with endemic exposure to Coccidioides spp. were tested more routinely for this
possibility. This is now the recommendation of the Arizona Department of Health
Services and a growing number of Arizona state medical specialty societies and
other professional organizations (Tsang et al., 2010).
With the commercial and recreational growth of the southwestern United
States, coccidioidomycosis has become an increasing problem for the rest of the
country as well. For example, persons who develop a respiratory illness within
a month after returning from vacation or business conferences in south-central
Arizona would have the same risk (approximately 30 percent) that their illness is
due to Coccidioides spp. as would residents of the endemic regions. Using Ari-
zona Department of Tourism statistics for 2008, the chance of an individual visi-
tor developing any clinical illness would be expected to be small (approximately
1 in 17,000). However, because more than 22 million persons visit Arizona for
an average of 4 to 5 days, the total number of illnesses occurring after leaving
Arizona would add up to more than 1,300 per year. Evidence suggests that most
of these illnesses would be diagnosed incorrectly in the course of routine medical
care (Standaert et al., 1995).
Even if physicians obtain appropriate testing, establishing a diagnosis of
early coccidioidal infection is often difficult. Coccidioidal serology is very spe-
cific when results are positive. Moreover, in progressive forms of infection, serol-
ogy is very likely to be diagnostic (Fish et al., 1990; Pappagianis and Zimmer,
1990). However, these tests are not nearly as sensitive early in the course of the
APPENDIX A 201
azole, and caspofungin (Gonzalez et al., 2001, 2007), which already have FDA
approval for other fungal diseases, there are very limited or no controlled clinical
trials conducted in patients with any form of coccidioidomycosis.
One exception to this pattern has been the persistent interest in bringing
nikkomycin Z into clinical trials. Nikkomycin Z is a competitive inhibitor of
chitin synthase, first discovered by German scientists in the 1970s. It was part of
a fungicide discovery program at the Bayer Company (Fiedler, 1988). Its poten-
tial as a therapeutic for coccidioidomycosis was identified in the 1980s (Hector
et al., 1990). In mice, nikkomycin Z treatment produced sterile lungs under
conditions in which the lungs of untreated mice yielded several million viable
fungal colonies. This observation raises the possibility that nikkomycin Z might
offer a curative treatment for coccidioidomycosis. If so, this would provide even
more incentive to diagnose coccidioidomycosis early in order to eradicate it and
thereby prevent later and serious complications. Clinical development of nikko-
mycin Z was begun in the 1990s by Shaman Pharmaceuticals (Galgiani, 2007).
However, the program became inactive when Shaman ceased to exist in 2000, and
for several years nikkomycin Z development remained dormant.
In 2005, the University of Arizona acquired the program along with several
kilograms of bulk nikkomycin Z that remained from Shaman’s program. Since
then faculty at the University of Arizona and a small start-up company, Valley
Fever Solutions, have successfully competed for research awards and small busi-
ness grants from the National Institutes of Health (NIH) and the FDA Office of
Orphan Products Development. With these funds as well as philanthropic support,
clinical trials with nikkomycin Z were resumed with a 2-week multidose safety
trial that was completed in 2009. This support is also being used to develop a
more efficient manufacturing process. Supplies of nikkomycin Z made by this
new process are planned to be available to begin a Phase II clinical trial in 2011 or
2012. It is hoped that this progress will advance the program sufficiently to attract
pharmaceutical or investment interest to complete the commercialization process.
the past 15 years, a collaboration of several research groups has yielded a number
of immunogenic coccidioidal proteins and vaccines prepared from recombinant
proteins with adjuvants, some of which have shown excellent protection in mice
and efficacy in primates (Cole et al., 2004; Cox and Magee, 2004; Herr et al.,
2007; Johnson et al., 2007; Shubitz et al., 2006; Tarcha et al., 2006).
The next step for existing recombinant vaccine candidates would be for them
to be moved into clinical trials. However, these candidates have met with major
challenges including developing a suitable manufacturing process; identifying a
suitable and available adjuvant; and compounding a suitable formulation appro-
priate for human experimentation (Galgiani, 2008). None of these challenges are
insurmountable, but all require significant development investment. The overall
impact of coccidioidomycosis within the endemic region is not so dissimilar to
that caused by polio in the United States before a polio vaccine was available
(approximately 10 per 100,000 population). However, the impact of coccidioido-
mycosis involves a much smaller population at risk as compared to the worldwide
distribution of polio. This difference in market size makes it unlikely that a com-
mercial vaccine manufacturer will invest in developing a coccidioidal vaccine
even though such a vaccine, once developed, could arguably be profitable to
manufacture and distribute (Barnato et al., 2001). Moving a coccidioidal vaccine
into clinical trials probably requires the discovery of new, more easily formulated
protective antigens; a breakthrough in vaccine technology that greatly reduces
the cost of development; or a growing public health imperative to underwrite the
costs needed for vaccine development.
Summary
Coccidioidomycosis is a major public health problem for a major, growing
segment of the U.S. population as well as other endemic regions throughout the
Western Hemisphere. A more complete understanding of its biology and ecology
where it exists in the endemic environment could lead to risk abatement strate-
gies not currently available. Improved recognition by healthcare professionals of
coccidioidomycosis as a cause of community-acquired pneumonia when it oc-
curs in their patients could improve management. This could be assisted further
by developing more sensitive and clinically available diagnostic tests based on
biosignatures such as DNA or proteins from the fungus itself. Curative therapies
are also needed, but none exist today. Finally, eliminating problems caused by
Coccidioides spp. might be possible if a preventive vaccine were developed. Even
though coccidioidomycosis is an orphan disease, pursuit of these objectives is
more than justified by the potential public health benefit and the reduced medical
costs to society that their achievement would provide.
204 FUNGAL DISEASES
Acknowledgments
This presentation was supported in part by Award Number U54AI065359
from the National Institute of Allergy and Infectious Diseases (NIAID). The
content is the sole responsibility of the authors and does not necessarily represent
the official views of the NIAID or NIH.
Disclosure
Dr. Galgiani is chief medical officer, chair of the board, and a significant
stock holder in Valley Fever Solutions, Inc.
References
Ampel, N. M., A. Giblin, J. P. Mourani, and J. N. Galgiani. 2009. Factors and outcomes associated
with the decision to treat primary pulmonary coccidioidomycosis. Clinical Infectious Diseases
48:172–178.
Barker, B. M., J. Tabor, L. Shubitz, R. Perill, and M. J. Orbach. 2010. Detection and phylogenetic
analysis of Coccidioides posadasii in Arizona soil samples. Fungal Ecology. In press.
Barnato, A. E., G. D. Sanders, and D. K. Owens. 2001. Cost-effectiveness of a potential vaccine for
Coccidioides immitis. Emerging Infectious Diseases 7:797–806.
Burt, A., B. M. Dechairo, G. L. Koenig, D. A. Carter, T. J. White, and J. W. Taylor. 1997. Molecular
markers reveal differentiation among isolates of Coccidioides immitis from California, Arizona
and Texas. Molecular Ecology 6:781–786.
Catanzaro, A., G. A. Cloud, D. A. Stevens, B. E. Levine, P. L. Williams, R. H. Johnson, A. Rendon,
L. F. Mirels, J. E. Lutz, M. Holloway, and J. N. Galgiani. 2007. Safety, tolerance, and efficacy
of posaconazole therapy in patients with nonmeningeal disseminated or chronic pulmonary
coccidioidomycosis. Clinical Infectious Diseases 45:562–568.
Chang, D. C., S. Anderson, K. Wannemuehler, D. M. Engelthaler, L. Erhart, R. H. Sunenshine, L. A.
Burwell, and B. J. Park. 2008. Testing for coccidioidomycosis among patients with community-
acquired pneumonia. Emerging Infectious Diseases 14:1053–1059.
Clark, K. A., and D. McAllister. 1996. Direct detection of Coccidioides immitis in clinical specimens
using target amplification. In Coccidioidomycosis, edited by H. E. Einstein and A. Catanzaro.
Proceedings of the Fifth International Conference. Washington, DC: National Foundation for
Infectious Diseases. Pp. 129–136.
Cole, G. T., J. M. Xue, C. N. Okeke, E. J. Tarcha, V. Basrur, R. A. Schaller, R. A. Herr, J. J. Yu, and
C. Y. Hung. 2004. A vaccine against coccidioidomycosis is justified and attainable. Medical
Mycology 42:189–216.
Comrie, A. C., and M. F. Glueck. 2007. Assessment of climate-coccidioidomycosis model: Model
sensitivity for assessing climatologic effects on the risk of acquiring coccidioidomycosis. Annals
of the New York Academy of Sciences 1111:83–95.
Converse, J. L., and R. E. Reed. 1966. Experimental epidemiology of coccidioidomycosis. Bacterio-
logical Reviews 30:678–695.
Cox, R. A., and D. M. Magee. 2004. Coccidioidomycosis: Host response and vaccine development.
Clinical Microbiology Reviews 17:804–839, table.
Crum-Cianflone, N. F. 2007. Coccidioidomycosis in the U.S. military: A review. Annals of the New
York Academy of Sciences 1111:112–121.
Dewsnup, D. H., J. N. Galgiani, J. R. Graybill, M. Diaz, A. Rendon, G. A. Cloud, and D. A. Stevens.
1996. Is it ever safe to stop azole therapy for Coccidioides immitis meningitis? Annals of Inter-
nal Medicine 124:305–310.
APPENDIX A 205
A9
17
51 *
FIGURE A9-1╇ Human infections with C. gattii, United States, December 2004–January
2011 (n = 79).
NOTE: * indicates patients that reported extensive travel to Washington and/or Oregon
A9-1 new
during the year before their illness onsets.
FIGURE A9-2╇ U.S. human cases of C. gattii, by year of illness onset (n = 71*).
Figure
* NOTE: Onset year is reported for A9-2.eps
62 patients and is estimated by initial report year for 9
bitmap2010 data are current as of January 2011;
patients for whom onset date was not available.
complete case data typically lag illness onset by several months.
PNW = Pacific Northwest.
Francisco in 1992 (Fraser et al., 2005). Taken together, these data suggest that
Southern California, but probably not the rest of the country, has long been an
endemic area for C. gattii. In addition, sporadic infections appear to be occurring
from other states, although the travel history of these patients and their potential
exposure site is unknown. Recently, environmental isolates of C. gattii have also
been found in Puerto Rico from a variety of cacti and tree material (Loperena-
Alvarez et al., 2010).
ized, and more likely to die from or with their infection than were B.C. patients
(Harris et al., 2010). The reasons for these differences are unclear, but might
include differences in case ascertainment in the PNW compared with B.C., which
has led to capture of a lower proportion of non-hospitalized C. gattii patients in
the PNW. Alternately, they could relate to the different genotypes seen in the
PNW compared with B.C. A comprehensive review of patient medical charts is
under way. It might help elucidate the differences and provide some insight into
whether these differences are real, and if so, why they exist.
The differences in U.S. C. gattii infections in the PNW, U.S. infections
outside the PNW, and infections occurring in other areas of the world might be a
function of C. gattii subtype, tropism, environmental distribution, or surveillance
bias. That is, C. gattii infections might not be reported completely from other
areas because they have not been looked for systematically elsewhere. Regardless
of the reasons for the outbreak in North America, reports of the outbreaks do not
appear to be exclusively due to temporal changes in surveillance. Retrospective
speciation studies of isolates from B.C. before 1999 (Fyfe et al., 2008) and from
the Seattle area before 2004 (Upton et al., 2007) suggested that the increase in
reported cases represents a true increase in infections in the region. It remains to
be seen whether the U.S. C. gattii infections outside of the PNW are sporadic,
travel associated, or linked to part of a larger emerging health issue in other areas
of the United States. Below is a discussion about efforts to conduct surveillance
for C. gattii infections outside of the PNW.
syndrome (IRIS) (Ecevit et al., 2006; Lui et al., 2006). IRIS is a paradoxical
clinical deterioration that is well documented during treatment of cryptococcosis
following initiation of antiretroviral therapy in AIDS patients (Woods et al., 1998)
and is thought to be due to an overzealous “rebound” immune response in the
presence of significant amounts of infecting pathogen. An IRIS-like syndrome
has also been documented in patients infected with C. gattii, where the syndrome
was suggested to be due to concomitant immune rebound and decreases in IL-10
(Einsiedel et al., 2004). These same factors may contribute to the severity of C.
gattii infection in immunocompetent hosts. At least one report exists of a patient
whose condition improved with steroid treatment, suggesting that an overly
functional immune system could confound treatment efforts in some patients
(Lane et al., 2004).
Discussion
The described outbreaks of C. gattii infection in the temperate climates of
B.C. and the PNW demonstrate a much less restrictive geographic range for C.
gattii than previously thought, and a broader range of persons who are suscep-
tible to infection. In particular, a compromised immune status now appears to be
a significant risk factor for at least some subtypes of C. gattii infection (CDC,
2010; Galanis and MacDougall, 2010). In addition, data from patients associated
with these outbreaks suggest that different C. gattii genotypes might infect dif-
ferent types of patients, and/or demonstrate different clinical courses resulting
from infection. The mere existence of an outbreak associated with C. gattii, never
previously reported, suggests that genetic components might be important for
pathogen spread in ways that are still poorly understood. More than ever, collect-
ing data is important that disease recognition and optimal treatment of C. gatiii
infections can be investigated. Several existing challenges now face the field.
Existing Challenges
One existing challenge is diagnosis of infection. Although several methods
exist to identify cryptococcal infections, including culture, India Ink stains of
cerebrospinal fluid or sputum (Cohen, 1984), and commercially available cryp-
tococcal antigen (CrAg) test kits (Saha et al., 2008), these methods cannot dis-
tinguish between C. neoformans and C. gattii. A simple way to confirm whether
or not a cryptococcal isolate is species gattii is to plate the isolate on canavanine-
glycine bromothymol blue (CGB) agar (Klein et al., 2009; Kwon-Chung et al.,
1982), where C. neoformans will leave the medium unaffected in color (yellow
to green) due to a failure to grow, and C. gattii will turn the medium blue due
to use of glycine as a carbon source. This medium is currently available from
at least one commercial supplier, but is not widely used in U.S. clinical micro-
biology labs. Thus, many C. gattii infections likely are being misdiagnosed as
APPENDIX A 219
C. neoformans. Ensuring that clinicians are aware of C. gattii infection and the
possible need for clinical differentiation from C. neoformans, and that their refer-
ence laboratories are able to speciate Cryptococcus isolates (and have an interest
in doing so), is critical to evaluate fully the geographic spread of disease and the
clinical spectrum of infections.
Investigating whether or not the most recent findings warrant modified treat-
ment guidelines is an additional challenge. The Infectious Diseases Society of
America published guidelines in 2010 (Perfect et al., 2010) that refer to differ-
ences in the treatment of C. gattii infections, compared with C. neoformans infec-
tions: specifically, C. gattii infections might require lengthier, more aggressive
treatment when compared with C. neoformans infections. The increased pro-
pensity for C. gattii to form cryptococcomas is also noted (Perfect et al., 2010).
However, the guidelines were largely based on data from C. gattii infections
occurring in Australia and Papua New Guinea. Increasingly, our data suggest
that even among C. gattii infections, not all cryptococcal infections are alike.
However, it is unclear which factors—infecting species, infecting subtype, host
immune status, or perhaps even host genetics—are most influential on patient
presentation and infection. Data from rigorous clinical studies are of utmost
importance in ensuring that clinician guidelines provide sufficient guidance to
optimize patient care. To this end, a large-scale, longitudinal chart review of
C. gattii infections is ongoing as a collaborative effort among Australia, B.C.,
and the United States, designed to address some of these questions. Results are
expected sometime in 2012.
Finally, the development of prevention messages is a challenge. Unlike C.
neoformans, which grows in pigeon feces, C. gattii appears to live in associa-
tion with trees and soil surrounding them (Springer and Chaturvedi, 2010). The
tree type appears to be less important than the presence of a wood substrate for
growth, and C. gattii has to date been associated with more than 50 tree species
(Randhawa et al., 2001; Springer and Chaturvedi, 2010). It has also been found
in air and water samples. These findings notwithstanding, C. gattii has not been
found ubiquitously around the globe in a distribution similar to C. neoformans,
and thus we can postulate that at least some environmental restrictions remain
in place for this organism. Environmental organisms present a specific challenge
for public health prevention because infections are usually relatively rare, and
difficult to avoid without draconian measures (e.g., staying indoors and purifying
air). This represents a quandary for public health officials. It remains to be seen
whether “hot spots” of infection exist in the PNW for which generalized recom-
mendations can be made that would benefit patient health, perhaps for subgroups
of higher risk patients. The benefits of outdoor activity would need to be weighed
against any risk calculated for these patients, and such recommendations are
bound to be controversial.
220 FUNGAL DISEASES
Conclusion
The increase in the number of reports during the past decade related to the
occurrence of C. gattii infections outside of traditional endemic tropical and
subtropical regions has provided excellent opportunities to learn more about this
important pathogen. The differences between individual cryptococcal infections
appear to be linked not only to patient immune status and infecting species, but
also to genetic subtypes within a species. It is unclear if the species and subtypes
have preferences for infection among certain patient types, possibly due to a need
for host immune support (or lack thereof) for replication, or if differences in
environmental colonization patterns might influence the type of patient infected.
For example, a pathogen with a ubiquitous distribution and a preference for im-
munocompromised patients will have a much higher infection rate and a much
higher immunocompromised to immunocompetent patient ratio than would a
pathogen with “hot spot distribution,” which would infect fewer patients overall,
but be limited to patients living in its area of environmental distribution (most of
APPENDIX A 221
References
Brandt, M. E., L. C. Hutwagner, L. A. Klug, W. S. Baughman, D. Rimland, E. A. Graviss, R. J.
Hamill, C. Thomas, P. G. Pappas, A. L. Reingold, and R. W. Pinner. 1996. Molecular subtype
distribution of Cryptococcus neoformans in four areas of the United States. Cryptococcal Dis-
ease Active Surveillance Group. Journal of Clinical Microbiology 34(4):912–917.
Bustamante Rufino, B., and D. Swinne. 1998. [Cryptococcus neoformans var. gattii isolates from two
Peruvian patients.]. Revista Iberoamericana de Micología 15(1):22–24.
Butler-Wu, S. M., and A. P. Limaye. 2011. A quick guide to the significance and laboratory identi-
fication of Cryptococcus gattii. American Society of Microbiology. http://www.asm.org/asm/
images/pdf/Clinical/cgattii.pdf (accessed March 28, 2011).
Byrnes, E. J., III, W. Li, Y. Lewit, J. R. Perfect, D. A. Carter, G. M. Cox, and J. Heitman. 2009. First
reported case of Cryptococcus gattii in the Southeastern USA: Implications for travel-associated
acquisition of an emerging pathogen. PLoS One 4(6):e5851.
Byrnes, E. J., III, W. Li, Y. Lewit, H. Ma, K. Voelz, P. Ren, D. A. Carter, V. Chaturvedi, R. J. Bildfell,
R. C. May, and J. Heitman. 2010a. Emergence and pathogenicity of highly virulent Cryptococ-
cus gattii genotypes in the northwest United States. PLoS Pathogens 6(4):e1000850.
———. 2010b. Examination of Cryptococcus gattii isolates from HIV/AIDS patients uncovers a
diverse population of VGIII molecular type isolates endemic in Southern California. Paper pre-
sented at Meeting of the Infectious Diseases Society of America, Vancouver, Canada, October
24, 2010.
Castanon-Olivares, L. R., R. Arreguin-Espinosa, G. Ruiz-Palacios y Santos, and R. Lopez-Martinez.
2000. Frequency of Cryptococcus species and varieties in Mexico and their comparison with
some Latin American countries. Revista Latinoamericana de Microbiologia 42(1):35–40.
CDC (Centers for Disease Control and Prevention). 2010. Emergence of Cryptococcus gattii—Pacific
Northwest, 2004–2010. Morbidity and Mortality Weekly Report 59(28):865–868.
222 FUNGAL DISEASES
Chaturvedi, S., M. Dyavaiah, R. A. Larsen, and V. Chaturvedi. 2005. Cryptococcus gattii in AIDS
patients, southern California. Emerging Infectious Diseases 11(11):1686–1692.
Chau, T. T., N. H. Mai, N. H. Phu, H. D. Nghia, L. V. Chuong, D. X. Sinh, V. A. Duong, P. T. Diep,
J. I. Campbell, S. Baker, T. T. Hien, D. G. Lalloo, J. J. Farrar, and J. N. Day. 2010. A prospec-
tive descriptive study of cryptococcal meningitis in HIV uninfected patients in Vietnam—high
prevalence of Cryptococcus neoformans var grubii in the absence of underlying disease. BMC
Infectious Diseases 10:199.
Chen, J., A. Varma, M. R. Diaz, A. P. Litvintseva, K. K. Wollenberg, and K. J. Kwon-Chung. 2008.
Cryptococcus neoformans strains and infection in apparently immunocompetent patients, China.
Emerging Infectious Diseases 14(5):755–762.
Chen, S., T. Sorrell, G. Nimmo, B. Speed, B. Currie, D. Ellis, D. Marriott, T. Pfeiffer, D. Parr, and
K. Byth. 2000. Epidemiology and host- and variety-dependent characteristics of infection due
to Cryptococcus neoformans in Australia and New Zealand. Australasian Cryptococcal Study
Group. Clinical Infectious Diseases 31(2):499–508.
Choi, Y. H., P. Ngamskulrungroj, A. Varma, E. Sionov, S. M. Hwang, F. Carriconde, W. Meyer,
A. P. Litvintseva, W. G. Lee, J. H. Shin, E. C. Kim, K. W. Lee, T. Y. Choi, Y. S. Lee, and K. J.
Kwon-Chung. 2010. Prevalence of the VNIc genotype of Cryptococcus neoformans in non-
HIV-associated cryptococcosis in the Republic of Korea. FEMS Yeast Research 10(6):769–778.
CIA (Central Intelligence Agency). 2010. The world factbook: United States. https://www.cia.gov/
library/publications/the-world-factbook/geos/us.html (accessed October 31, 2010).
Cohen, J. 1984. Comparison of the sensitivity of three methods for the rapid identification of Cryp-
tococcus neoformans. Journal of Clinical Pathology 37(3):332–334.
Diaz, M. R., T. Boekhout, B. Theelen, and J. W. Fell. 2000. Molecular sequence analyses of the
intergenic spacer (IGS) associated with rDNA of the two varieties of the pathogenic yeast,
Cryptococcus neoformans. Systematic and Applied Microbiology 23(4):535–545.
Discover Magazine. 2010. A tropical, fatal fungus gains a foothold in the Pacific Northwest. http://
blogs.discovermagazine.com/80beats/2010/04/23/a-tropical-fatal-fungus-gains-a-foothold-in-
the-Pacific-Northwest/ (accessed October 25, 2010).
Duncan, C., H. Schwantje, C. Stephen, J. Campbell, and K. Bartlett. 2006. Cryptococcus gat-
tii in wildlife of Vancouver Island, British Columbia, Canada. Journal of Wildlife Diseases
42(1):175–178.
Dwyer, V. 2003. ClinMicroNet—Sharing experiences and building knowledge virtually. Clinical
Microbiology Newsletter 25(16):121–125.
Ecevit, I. Z., C. J. Clancy, I. M. Schmalfuss, and M. H. Nguyen. 2006. The poor prognosis of central
nervous system cryptococcosis among nonimmunosuppressed patients: A call for better dis-
ease recognition and evaluation of adjuncts to antifungal therapy. Clinical Infectious Diseases
42(10):1443–1447.
Einsiedel, L., D. L. Gordon, and J. R. Dyer. 2004. Paradoxical inflammatory reaction during treatment
of Cryptococcus neoformans var. gattii meningitis in an HIV-seronegative woman. Clinical
Infectious Diseases 39(8):e78–e82.
Ellis, D. H. 1987. Cryptococcus neoformans var gattii in Australia. Journal of Clinical Microbiology
25(2):430–431.
Escandon, P., A. Sanchez, M. Martinez, W. Meyer, and E. Castaneda. 2006. Molecular epidemiology
of clinical and environmental isolates of the Cryptococcus neoformans species complex reveals
a high genetic diversity and the presence of the molecular type VGII mating type a in Colombia.
FEMS Yeast Research 6(4):625–635.
Fraser, J. A., S. S. Giles, E. C. Wenink, S. G. Geunes-Boyer, J. R. Wright, S. Diezmann, A. Allen, J. E.
Stajich, F. S. Dietrich, J. R. Perfect, and J. Heitman. 2005. Same-sex mating and the origin of
the Vancouver Island Cryptococcus gattii outbreak. Nature 437(7063):1360–1364.
Fyfe, M., W. Black, M. Romney, et al. 2002. Unprecedented outbreak of Cryptococcus neoformans
var. gattii infections in British Columbia, Canada. Paper presented at the Fifth International
Conference on Cryptococcus and Cryptococcosis, Adelaide, Australia, March 3–7.
APPENDIX A 223
Fyfe, M., L. MacDougall, M. Romney, M. Starr, M. Pearce, S. Mak, S. Mithani, and P. Kibsey. 2008.
Cryptococcus gattii infections on Vancouver Island, British Columbia, Canada: Emergence of a
tropical fungus in a temperate environment. Canada Communicable Disease Report 34(6):1–12.
Galanis, E., and L. MacDougall. 2010. Epidemiology of Cryptococcus gattii, British Columbia,
Canada, 1999–2007. Emerging Infectious Diseases 16(2):251–257.
Gatti, F., and R. Eeckels. 1970. An atypical strain of Cryptococcus neoformans (San Felice) Vuillemin
1894. Description of the disease and of the strain. Annales des Sociétés Belges de Médecine
Tropicale, de Parasitologie, et de Mycologie 50(6):689–693.
Harris, J., S. R. Lockhart, N. Marsden-Haug, R. Wohrle, C. Free, E. DeBess, and T. Chiller. 2010.
Poster 642. Cryptococcus gattii: Emergence of a novel pathogen in the United States Pacific
Northwest. Paper presented at the Infectious Diseases Society of America Meeting, Vancouver,
Canada, October 20–24, 2010.
Hoang, L. M., J. A. Maguire, P. Doyle, M. Fyfe, and D. L. Roscoe. 2004. Cryptococcus neoformans
infections at Vancouver Hospital and Health Sciences Centre (1997–2002): Epidemiology, mi-
crobiology and histopathology. Journal of Medical Microbiology 53(Pt 9):935–940.
Hutchison, C. 2010. Fatal fungus Cryptococcus gattii: Experts say fears overblown. http://abcnews.
go.com/Health/Wellness/fatal-fungus-sparks-fear-worry/story?id=10438475 (accessed October
5, 2010).
Iqbal, N., E. E. DeBess, R. Wohrle, B. Sun, R. J. Nett, A. M. Ahlquist, T. Chiller, and S. R. Lockhart.
2010. Correlation of genotype and in vitro susceptibilities of Cryptococcus gattii strains from
the Pacific Northwest of the United States. Journal of Clinical Microbiology 48(2):539–544.
Kidd, S. E., P. J. Bach, A. O. Hingston, S. Mak, Y. Chow, L. MacDougall, J. W. Kronstad, and K. H.
Bartlett. 2007. Cryptococcus gattii dispersal mechanisms, British Columbia, Canada. Emerging
Infectious Diseases 13(1):51–57.
Klein, K. R., L. Hall, S. M. Deml, J. M. Rysavy, S. L. Wohlfiel, and N. L. Wengenack. 2009. Identi-
fication of Cryptococcus gattii by use of L-canavanine glycine bromothymol blue medium and
DNA sequencing. Journal of Clinical Microbiology 47(11):3669–3672.
Kwon-Chung, K. J., and J. E. Bennett. 1984a. Epidemiologic differences between the two varieties of
Cryptococcus neoformans. American Journal of Epidemiology 120(1):123–130.
———. 1984b. High prevalence of Cryptococcus neoformans var. gattii in tropical and subtropi-
cal regions. Zentralblatt Fuer Bakteriologie,Microbiologie, und Hygiene (Reihe A) 257(2):
213–218.
Kwon-Chung, K. J., I. Polacheck, and J. E. Bennett. 1982. Improved diagnostic medium for sepa-
ration of Cryptococcus neoformans var. neoformans (serotypes A and D) and Cryptococcus
neoformans var. gattii (serotypes B and C). Journal of Clinical Microbiology 15(3):535–537.
Kwong-Chung, K., T. Boekhout, J. W. Fell, and M. Diaz. 2002. Proposal to conserve the name Cryp-
tococcus gattii against C. hondurianus and C. bacillisporus (Basidiomycota, Hymenomycetes,
Tremellomycetidae). Taxon 51:804–806.
Lalloo, D., D. Fisher, S. Naraqi, I. Laurenson, P. Temu, A. Sinha, A. Saweri, and B. Mavo. 1994.
Cryptococcal meningitis (C. neoformans var. gattii) leading to blindness in previously healthy
Melanesian adults in Papua New Guinea. Quarterly Journal of Medicine 87(6):343–349.
Lane, M., J. McBride, and J. Archer. 2004. Steroid responsive late deterioration in Cryptococcus
neoformans variety gattii meningitis. Neurology 63(4):713–714.
Laurenson, I., S. Naraqi, N. Howcroft, I. Burrows, and S. Saulei. 1993. Cryptococcal meningitis
in Papua New Guinea: Ecology and the role of eucalypts. Medical Journal of Australia 158
(3):213.
Laurenson, I. F., A. J. Trevett, D. G. Lalloo, N. Nwokolo, S. Naraqi, J. Black, N. Tefurani, A. Saweri,
B. Mavo, J. Igo, and D. A. Warrell. 1996. Meningitis caused by Cryptococcus neoformans var.
gattii and var. neoformans in Papua New Guinea. Transactions of the Royal Society of Tropical
Medicine and Hygiene 90(1):57–60.
224 FUNGAL DISEASES
A10
Abstract
How microbial pathogens emerge to cause outbreaks and become established
as agents of disease in humans involves genetic exchange, zoonotic transmission,
and perturbations of ecosystems and habitats. The threat of emerging infec-
tious diseases is particularly poignant for eukaryotic pathogens, the fungi, and
parasites, given that these microbes are more difficult to treat and have complex
genomes and lifecycles. A sobering recent development has been the emergence
and reemergence of several fungal pathogens in both humans and other animals,
including Geomyces destructans in bats, Batrachochytrium dendrobatidis in
amphibians, Nosema ceranae in bees (colony collapse disorder), and Cryptococ-
cus gattii in humans and other animals in the Pacific Northwest. Here we review
issues surrounding the C. gattii outbreak that began on Vancouver Island in 1999
and has expanded into the United States in Washington, Oregon, and California
and has the potential to expand further. The focus will be on the emergence of C.
gattii in the United States, including the appearance of a novel, highly virulent
genotype and the potential role of sexual reproduction in the emergence of novel
pathogens and their dispersal via airborne spores.
Introduction
The early history of cryptococcosis was documented in single or small se-
ries of cases. From an initial case of tibial osteomyelitis with the encapsulated
Cryptococcus neoformans yeast in 1895 until a seminal monograph on this
disease by Littman and Zimmer in 1956, the entire repertoire of reports in the
medical literature numbered less than 300 cases (Littman and Zimmer, 1956).
This was a humble beginning for this cosmopolitan, encapsulated basidiomycete
that has now emerged into an outbreak mode in the new millennium. In the first
half-century of its known existence, many of the clinical features of cryptococ-
cosis were well described, including its propensity to invade the central nervous
Vancouver Island C. gattii Outbreak and Expansion into the United States
Our focus in this chapter will be the C. gattii outbreak that began on Van-
couver Island in 1999 and has now expanded into the Canadian mainland in
British Columbia and into the United States. A considerable body of knowledge
is available with respect to the life and virulence cycles for this pathogenic
yeast (Heitman, 2011). C. neoformans and C. gattii are closely aligned species.
C. neoformans is prevalent in the environment globally, and C. gattii has been
FIGURE A10-1╇The C. gattii outbreak expanded into, and emerged within, the United
Figure A10-1.eps
States (n = 56). BC = British Columbia.
bitmap
ing multilocus sequence typing (MLST) (Bovers et al., 2008; Fraser et al., 2005).
This distinction of cryptic species is important because the VGII cryptic species
is causing the outbreak in the Pacific Northwest, whereas the VGI lineage is more
prevalent and more commonly causing disease in Australia (see Heitman, 2011).
The outbreak of C. gattii in the Pacific Northwest began in 1999 and prior
to this, no C. gattii infections were reported as causing disease in this region of
the world. The first cases came to the attention of astute clinicians and veterinar-
ians on the southeastern shores in Naniamo and Parksville on Vancouver Island
(Duncan et al., 2005; Hoang et al., 2004; Stephen et al., 2002). Over the past
decade, there have been approximately 260 cases with an approximately 10 per-
cent attributable mortality rate, and many infections in animals. Interested read-
230 FUNGAL DISEASES
ers are referred to a reprint36 included in this volume from Karen Bartlett, who
played a critical role in identifying the environmental source of the organism on
Vancouver Island (Bartlett et al., 2008), and to the manuscript from Julie Harris
from the CDC Cryptococcus working group.37 Our charge in this article was to
consider two aspects of the outbreak: first, its expansion into the United States,
and second, the possible roles of sexual reproduction in the origin of the outbreak
isolates and the ongoing production of airborne infectious spores.
The first cases associated with the Vancouver Island outbreak in patients
from mainland British Columbia with no travel history to the island appeared
in 2003–2004 (Bartlett et al., 2008). Environmental sampling studies provided
evidence that C. gattii expanded across the water to a broader niche, including
the Canadian mainland (MacDougall et al., 2007). The southern tip of Vancouver
Island is very close to the U.S. border, and therefore a key question was if and
when the outbreak might spread into the United States. The San Juan Archipelago
is a part of Washington state, located as near as 5 km from the gulf islands off
the coast of Vancouver Island. The first C. gattii index case in the United States
was an elderly patient with leukemia on Orcas Island, Washington, who presented
with a pulmonary C. gattii nodule in January 2006 (Upton et al., 2007). Since
then, the outbreak has expanded into the United States from Canada, and we
summarize here the evidence and key findings.
From 2006 to 2010, approximately 70 cases in patients have been reported,
and the most complete records are attributable to the efforts of the CDC Cryp-
tococcus working group. The cases in humans and animals are shown in Figure
A10-1, and the icons represent cases for which we have isolates in the laboratory
that have been molecularly analyzed. Two types of isolates are circulating on
Vancouver Island: the VGIIa/major genotype, which causes approximately 95
percent of the infections, and the VGIIb/minor genotype, which represents the
remaining 5 percent of isolates (Fraser et al., 2005). Both of these genotypes have
been found in the environment on Vancouver Island and cause infection in both
patients and animals, and both have now emerged within the United States. Of
particular interest is a completely novel genotype that has emerged in Oregon:
VGIIc, or the novel genotype (Byrnes et al., 2009, 2010). VGIIc has not yet been
found in Washington state, Vancouver Island, or anywhere else in the world; thus,
it is a completely new emergence from 2005 to 2010 in Oregon.
We can identify genotypes by applying MLST of genetic barcodes through-
out the genome. In this technique, coding and non-coding regions of genes are
PCR amplified, sequenced, and assigned a unique allele number. The alleles are
then color coded and organized in a tabular format in which each line represents
a different isolate associated with the outbreak or a global isolate in the strain
collection. This allows one to appreciate that there has been what appears to be a
large clonal expansion of the VGIIa/major genotype in the region that dominates
36╛╛See contributed manuscript by Bartlett in Appendix A (pages 101–116).
37╛╛See contributed manuscript by Harris in Appendix A (pages 207–225).
232 FUNGAL DISEASES
the outbreak on Vancouver Island and its expansion into Puget Sound and beyond
in Washington state. There are fewer isolates of the VGIIb/minor genotype, but
these have been found on Vancouver Island, in Oregon, and most recently in
Washington state. The VGIIc/novel genotype has thus far been found only in
Oregon and has several, unique alleles not found to date in any other isolates
examined from global sources. The CDC has identified one isolate from a patient
in Idaho that appears to be closely related to the VGIIc/novel genotype, but it
may be the result of travel exposure (DeBess et al., 2010).
Where did these VGII outbreak genotypes originate? The VGIIa/major geno-
type is indistinguishable across 30 MLST loci and several variable number of
tandem repeat loci from the NIH444 strain, which was isolated in the early 1970s
from a sputum sample from a patient in Seattle (Fraser et al., 2005). This isolate is
the type strain for C. gattii and is molecularly indistinguishable from the VGIIa/
major outbreak strain at all loci examined thus far. Therefore, the major outbreak
strain appears to have been in this geographic region for at least 40 years and
possibly even longer. The VGIIb/minor genotype isolates are indistinguishable
from isolates from a fully recombining sexual population in Australia (Campbell
et al., 2005a,b; Fraser et al., 2005). The VGIIc/novel genotype has thus far only
been identified in isolates collected in Oregon, and we have not observed it in a
large collection of more than 200 C. gattii isolates collected globally. It appears
as if the VGIIc/novel genotype either emerged locally in Oregon or, alternatively,
it may be present in an undersampled environmental niche that remains to be
discovered. Further evidence that this new genotype in Oregon is novel involves
haplotype network mapping. In this approach, the MLST alleles are compared
and we apply a computer algorithm to predict the ancestral allele. The alleles are
then organized into a diagram rooted with the ancestral allele, and alleles derived
from it by mutations and genetic drift are indicated with lines and circles. In this
analysis alleles that are arisen from the ancestral allele long ago lie closest to the
ancestral allele, whereas alleles that arose recently lie distal. The MLST alleles
that are private to the VGIIc genotype are distal in these haplotype networks,
suggesting that they arose recently. The alleles that the VGIIc alleles are derived
from come from isolates that originated from either Australia or South America
(Byrnes et al., 2010). This analysis then suggests that those alleles might repre-
sent possible sites of origin of at least some of the genetic material in the Oregon
VGIIc/novel outbreak isolate.
We have conducted mammalian virulence studies in a mouse inhalation
model, both at the Wadsworth Center in Albany with our collaborators Ping Ren,
Sudha Chaturvedi, and Vishnu Chaturvedi and also at Duke University (Byrnes
et al., 2010; Fraser et al., 2005). The VGIIa/major genotype and the VGIIc/novel
genotype isolates are both highly virulent in the mouse model used, whereas the
minor VGIIb genotype is considerably attenuated by direct comparison (Byrnes
et al., 2010). Isolates collected globally that share many but not all markers with
the VGIIa/major genotype are considerably less virulent than VGIIa (Byrnes
APPENDIX A 233
et al., 2010). This includes, for example, isolates from both California and South
America. Of note, the NIH444 type strain (a VGIIa/major genotype isolate) is
also somewhat less virulent than contemporary outbreak VGIIa/major genotype
strains, which may reflect either attenuation as a result of long-term storage and
passage; variation in virulence even among the VGIIa/major isolates; or unknown
genetic differences between NIH444 and the VGIIa/major outbreak genotypes
in regions of the genome not yet analyzed. Therefore, it appears as if virulence
has increased with the emergence of the VGIIa/major outbreak genotype and the
VGIIc/novel genotype from their original source strains.
Several investigators have been addressing why outbreak isolates are more
virulent than other genotypes. There have been two prescient studies. Jim Kro-
nstad and colleagues at the University of British Columbia in Vancouver have
shown that there is less protective inflammation in the lungs of mice infected
with the outbreak isolates (Cheng et al., 2009). This may then lead to increased
virulence if the host fails to mount a sufficient, protective immune response. In
addition, the work from Robin May’s lab at the University of Birmingham has
described an increased intracellular proliferation rate of the VGIIa and VGIIc
outbreak isolates in macrophages and linked mitochondrial function to the robust-
ness of their interaction with host immune cells (Byrnes et al., 2010; Ma et al.,
2009). Given that Cryptococcus is a pathogen able to grow either outside or
inside of host immune cells, an enhanced proliferation rate in the context of the
macrophage intracellular milieu may lead to enhanced virulence. These studies
are excellent starting points to use to begin to dissect the virulence mechanisms
at the interface with host immune cells in the lung.
To summarize, this outbreak was originally restricted geographically to
Vancouver Island and then spread to the mainland of British Columbia. Starting
in 2006 (and maybe earlier in one case in Oregon in 2005), it emerged within the
United States. Thus, there clearly appears to have been a geographic expansion
from Vancouver Island across Puget Sound to reach Washington, mostly involv-
ing the VGIIa/major genotype. However, there is more diversity in strains from
Oregon, involving both the VGIIa/major and VGIIb/minor genotypes observed
on Vancouver Island, but also including the novel VGIIc genotype that has not
been identified on Vancouver Island. In essence, this looks like an outbreak
within an outbreak, which may be of independent origins. It is as though two
pebbles have been dropped into a pond at different times, one earlier than the
other, and they have generated concentric waves that are now expanding outward
and intersecting. There is considerable concern that this outbreak will continue
to expand geographically given that there are cases and isolates now from Idaho
and California (Personal communication, Shawn Lockhart, CDC Mycotic Dis-
eases Branch, May 2011). There is also a well-documented risk to travelers to the
endemic region, who then return to locations around the world with an infection
of the VGIIa/major isolate and present with a very unusual fungal infection that
is not commonly described in those regions and that might confuse clinicians
234 FUNGAL DISEASES
(Chambers et al., 2008; Georgi et al., 2009; Hagen et al., 2010; Lindberg et al.,
2007).
and Heitman, 2002). This fact raised a central conundrum for the entire field of
cryptococcal pathogenesis: If there is a well-maintained a-a opposite-sex sexual
cycle, but the a mating type is extremely rare, if not completely absent in many
populations, how then can a sexual cycle be an important step of the infectious
cycle for this fungus? It might be that it is not important, such that these unisexual
populations were just clonally reproducing mitotically as yeasts. However, we
discovered that C. neoformans can undergo an unusual sexual cycle involving
only one of the two mating types (Lin et al., 2005).
The opposite-sex cycle that we have known about for 30 years is depicted in
the lower panel of Figure A10-3. Then in 2005, an alternative sexual cycle was
reported called unisexual reproduction or same-sex mating, and this is depicted
in the upper panel of Figure A10-3 (Lin et al., 2005; Wang and Lin, 2011). This
alternative sexual cycle only involves a cells. They can fuse with another a cell
from the population, or they can, in extreme cases, fuse with themselves, undergo
meiosis, and produce spores.
You might wonder, what would be the point of undergoing sexual reproduc-
tion with yourself? There is no genetic diversity to admix in this circumstance.
We always think about sexual reproduction as involving two parents with very
different genetic compositions. It turns out that mating with yourself can also
lead to the generation of genetic diversity. Sex itself can serve as something of
a mutagen to generate genotypic and phenotypic diversity. This strategy turns
completely on its head what we traditionally think of as the primary role of sexual
reproduction. An analogy might be the appearance of mismatch repair muta-
tor mutations in bacterial pathogens, which arise to result in the generation of
genome-wide mutations and are then swept from the population as a consequence
of their concomitant deleterious effects.
As an example, in preliminary studies we isolated 100 progeny produced
by this same-sex mating cycle and looked for phenotypic diversity. We have
found within this set of just 100 isolates clear examples of azole resistance,
temperature-sensitive growth, hyperfilamentous growth, and increased melanin
production. By comparison, analysis of 100 progeny produced by mitotic clonal
growth yielded no such examples. Many of the phenotypically distinct isolates
turn out to be aneuploid in one way or another. This is based on comparative ge-
nome hybridization analysis showing that one of these morphologically distinct
isolates now has an extra copy of chromosome 10. The presence of an extra copy
of a chromosome is termed aneuploidy. We often associate aneuploidy with very
deleterious consequences. We do not have to go further than Down syndrome
as an example to think about what might be bad about aneuploidy. On the other
hand, in the microbial kingdom, aneuploidy can be a rich source of phenotypic
diversity. There are well-documented studies in Candida albicans that an a spe-
cial type of chromosome formed by duplicated arms of a chromosome (termed
an isochromosome) can form and drive fluconazole resistance (Selmecki et al.,
2006, 2008, 2009). Similarly, we also know that azole heteroresistance in Crypto-
coccus is driven by aneuploidy for chromosome 1 (Hu et al., 2008; Sionov et al.,
2009, 2010). A variety of other experimental evolution studies have brought to
the forefront the idea that aneuploidy might be a driving force for genotypic and
phenotypic plasticity (Pavelka et al., 2010; Rancati et al., 2008; Torres et al.,
2007, 2010). So this may be one mechanism by which a same-sex mating cycle
could engender phenotypic and genotypic diversity in a population without need-
ing a partner that is genetically divergent.
Another way to view same-sex mating is that it is a mechanism for selfing.
Selfing/inbreeding is common in fungi and the paradigm is mating type switching
in S. cerevisiae, which allows mother–daughter cell mating. Same-sex mating is
another route to self, but not via mating type switching. It is a question at some
level of outcrossing vs. inbreeding. Same-sex mating superimposes much more
limited genetic diversity on a well-adapted fungal genome that has run the gaunt-
let of natural selection. As such, unisexual reproduction may confer one of the
benefits of sex (generation of diversity), but at the same time avoid one of its costs
(breaking apart well-adapted genomic/genetic configurations). For organisms that
are well adapted to a niche, this more limited genetic diversity may better answer
subtle changes in selection pressures.
APPENDIX A 237
We have alluded to the fact that there are restricted geographic areas where
fungal sex is extant. Why is that? Why would you have these little areas and
pockets in Africa and that is where they have sex, and in the rest of the world they
don’t have sex except for the unusual unisexual cycle? At least for Cryptococcus,
what it looks like is, where they have opposite-sex mating seems to be where
there are restricted populations where both mating types still exist. We know the
most about this from C. neoformans, where there is an extant population that is
on a very specific indigenous tree in South Africa, in Botswana (Litvintseva et al.,
2011). But everywhere else in the world, you find just the alpha mating type (Hull
and Heitman, 2002). So there is a niche in nature where opposite-sex mating is
occurring. Until 5 years ago everyone presumed everywhere else it was asexual
and mitotic. But what we are coming to appreciate is that there is just a different
version of the sexual cycle occurring in these other populations where there is
just one mating type.
This is interesting as another classic example is Phytophthora infestans. In
Mexico and in Peru and other areas of South America, that is where the sexual
cycle occurs, new versions arise, and then they are often exported on potato
crops. That was the source of the Irish potato famine. But why there is a sexual
cycle in such a restricted place is not really clear. Toxoplasma gondii might be
another interesting pathogen to consider. It only has its sexual cycle in cats and
other felids, not in other animals of any sort (Heitman, 2006). So what is it about
the gastrointestinal tract of a cat that is so different from a mouse or a human that
allows this parasite to undergo its sexual cycle there? The parasite sexual cycles
are even more bizarre than some of the fungal ones, and restricted to extreme
niches (Heitman, 2006, 2010).
What about the specific example of C. gattii sexual reproduction? June
Kwon-Chung elucidated the C. gattii sexual cycle more than 30 years ago
(Kwon-Chung, 1976a,b). Spores were produced in her classic studies via an a-a
opposite-sex mating cycle that she defined. We were approached by several of
the Australian groups working on the VGI genotypes in the late 1990s, including
Wieland Meyer and Dee Carter, and we spent more than 5 years recapitulating
this sexual cycle under laboratory conditions. This required a great deal of heroic
effort on the part of several individuals in various laboratories, in large part be-
cause fecundity can be reduced with long-term passage and storage and because
the VGI molecular type that is predominant in Australia is more recalcitrant in
mating assays compared to the VGII and VGIII C. gattii molecular types. We
were finally able to recapitulate a sexual cycle for C. gattii involving the for-
mation of dikaryotic hyphae with special cells linking the hyphal cells (fused
clamp cells) culminating in the production of basidia and basidiospores, all of
the morphological features of the sexual cycle, including meiotic recombination
(Campbell et al., 2005a,b; Fraser et al., 2003). Based on this advance, we were
able to show that the vast majority of the outbreak isolates from Vancouver Is-
land are fertile in laboratory crosses (Byrnes et al., 2010; Campbell et al., 2005a;
238 FUNGAL DISEASES
Fraser et al., 2003). These are a-a matings that were conducted, but we stress
the point that every single isolate that is associated with the outbreak is of the a
mating type. No one has found any a isolates occurring anywhere on Vancouver
Island or in Washington or Oregon.
How might mating occur on Vancouver Island and in the Pacific Northwest in
this unisexual population? In previous studies we observed that an association of
Cryptococcus with plants can stimulate the fungal sexual cycle (Xue et al., 2007).
Part of the idea in testing for this phenomenon in the first place was that plants,
and more specifically trees, are the environmental niche for C. gattii (Heitman
et al., 2011). For the plant pathogenic fungus Ustilago maydis, which infects corn,
the infectious form is the filamentous dikaryotic hyphae produced by the sexual
cycle, which is stimulated by interaction with the plant. In fact, U. maydis has to
be in its sexual form to infect the plant host. We found something very similar
with Cryptococcus. We were unable to infect a plant efficiently with the haploid
yeast, but the filamentous dikaryotic state will infect plants and stimulate a plant
defense response (Xue et al., 2007). In turn, if we put seedlings distant from a
fungal mating mixture on a plate, the plants secrete small molecules, such as ino-
sitol, that stimulate completion of the sexual cycle and production of spores (Xue
et al., 2007, 2010). We observed this with seedlings of Arabidopsis or Eucalyptus
and are now exploring this with Douglas fir because of the link to this indigenous
tree species in the Pacific Northwest outbreak. This line of investigation suggests
that this may be one niche in nature where the sexual cycle is occurring to pro-
duce spores as small airborne infectious propagules.
Another critical line of investigation involves population genetic studies con-
ducted by Dee Carter from the University of Sydney, Australia. She has focused
on the sexual cycle that may be occurring in Eucalyptus tree hollows. She discov-
ered that in populations with both mating types, they can engage in opposite-sex
mating, but in other tree hollows where only a isolates are found, they are also
sexually recombining (Saul et al., 2008). So it seems as if the organism has two
extant sexual cycles, one opposite sex and one same sex, and whom you mate
with depends on who your neighbors are in the population. All of her studies
have focused on the VGI cryptic species, which has very limited genetic diversity
and an extremely high level of apparent inbreeding (Campbell and Carter, 2006;
Campbell et al., 2005a,b; Carter et al., 2007, 2011). All of the isolates on Vancou-
ver Island associated with the clonal outbreak are of a different cryptic species,
VGII. However, based on Dee Carter’s studies of VGI sexual reproduction, the
presumption is that there may be both forms of sexual reproduction occurring also
for the VGII isolates in nature, associated with the environmental niche in trees
or plants for the VGII lineage and involving both opposite and same-sex mating
depending on where a isolates are found in nature.
We have advanced the population genetic analysis in two ways with our
global isolates to look for measures of sexual reproduction and recombination
occurring in the population. One of these is a simple test called an allele com-
APPENDIX A 239
patibility test. The basic premise is that if you cross two strains that differ at
two markers, you obtain four types of progeny produced by sexual processes by
reassorting two unlinked loci (i.e., AB by ab yields AB, ab, Ab, and aB) (Carter
et al., 2011), which we all know from studying Punnett squares. So the multilo-
cus sequence loci can be analyzed as two unlinked alleles in the population. If
you find all four possible combinations, this is indicative of sexual reproduction
and recombination occurring in the population. By looking at the multilocus
sequence markers, we find that these measures of recombination are rampant
throughout the VGII global population (Byrnes et al., 2010). In addition, if we
just examine the sequences of the multilocus sequence alleles themselves, which
represent approximately 1 kilobase of sequence each, we can find examples of
hybrid recombinant alleles that implicate isolates as potential parents or potential
offspring (Byrnes et al., 2010). A lot of mitotic recombination must be occur-
ring to see evidence of recombination within just a random 1-kilobase sequence
in a 20-megabase genome. The potential parents that are identified by this type
of analysis originate from South America, Africa, and Australia (Byrnes et al.,
2010). These findings suggest that sexual reproduction is occurring globally in
multiple environments and locations.
We went on to examine these global isolates for fertility in the lab and
documented, by scanning with electron microscopy, the production of meiotic
spores, which are the products of sex that may be infectious propagules found in
the air on Vancouver Island. The majority of the a strains we analyzed are fertile.
We have found a very limited number of a strains in the global population. For
example, we found 6 out of 200 (~3 percent) are this minor mating type. Five
out of six of these a strains are fertile in the laboratory, and thus they may be
undergoing opposite-sex mating in nature. Because five of these six a isolates are
from South America, it may be a site in which opposite-sex mating is still extant
in the population (Escandon et al., 2007; Fraser et al., 2005; Ngamskulrungroj
et al., 2008). We know that C. neoformans has also retained extant opposite-sex
mating, but that it is geographically restricted to Botswana and South Africa and
occurring in the VNB C. neoformans lineage (Litvintseva et al., 2003, 2007).
When we return to consider the VGIIb minor genotype and its relationship
to global isolates, this provides critical insight. For instance, we looked at 30
MLST alleles, the VGIIb outbreak isolates are completely indistinguishable from
isolates from a fertile, unisexual, sexually recombining population in Australia
that was identified by Dee Carter and colleagues (Campbell et al., 2005a,b). It
suggests that this population may be the ancestral source for the VGIIb/minor
outbreak genotype isolates now circulating in the Pacific Northwest. While one
might posit just the opposite (transfer from Vancouver Island to Australia), the
diversity of the population in Australia supports this as the ancestral rather than
the derived population. Thus, the most parsimonious model is that the isolates
from the outbreak originated from Australia (Figure A10-4).
Based on a broad comparison of the isolates on Vancouver Island, the VGIIa/
240 FUNGAL DISEASES
αxα
αxα
Eucalyptus trees,
original same-sex
mating a x α?
αxα
Avirulent parent
in a fertile recombining population
=VGIIa/major, highly virulent
Acknowledgments
We thank Wenjun Li and Yonathan Lewit for their contributions to the analy-
sis of the outbreak; Dee Carter, Vishnu Chaturvedi, Jim Kronstad, Kieren Marr,
APPENDIX A 243
and Robin May for collaboration; Rory Duncan, Chris Lambros, and Dennis
Dixon from the National Institute of Allergy and Infectious Diseases (NIAID)
and Victoria McGovern from the Burroughs Wellcome Fund for their support;
and Stephen Johnston for his prescient questions. We also explicitly thank all of
the Forum members and the Institute of Medicine for shining a very bright light
on the fungal kingdom, and for the invitation to participate. Our research is sup-
ported by NIH/NIAID R37 grant AI39115.
References
Baro, T., J. M. Torres-Rodriguez, M. H. De Mendoza, Y. Morera, and C. Alia. 1998. First identifica-
tion of autochthonous Cryptococcus neoformans var. gattii isloated from goats with predomi-
nantly severe pulmonary disease in Spain. Journal of Clinical Microbiology 36(2):458–461.
Bartlett, K. 2010. Knowing where to look—environmental sources of cryptococcal disease in human
and animal residents in the Pacific Northwest. Presentation given at the December 14–15, 2010
public workshop, “Fungal Diseases: An Emerging Challenge to Human, Animal, and Plant
Health.” Forum on Microbial Threats, Institute of Medicine, Washington, DC.
Bartlett, K. H., S. E. Kidd, and J. W. Kronstad. 2008. The emergence of Cryptococcus gattii in British
Columbia and the Pacific Northwest. Current Infectious Disease Reports 10(1):58–65.
Botts, M. R., and C. M. Hull. 2010. Dueling in the lung: How Cryptococcus spores race the host for
survival. Current Opinion in Microbiology 13(4):437–442.
Botts, M. R., S. S. Giles, M. A. Gates, T. R. Kozel, and C. M. Hull. 2009. Isolation and characteriza-
tion of Cryptococcus neoformans spores reveal a critical role for capsule biosynthesis genes in
spore biogenesis. Eukaryotic Cell 8:595–605.
Bovers, M., F. Hagen, E. E. Kuramae, and T. Boekhout. 2008. Six monophyletic lineages identified
within Cryptococcus neoformans and Cryptococcus gattii by multi-locus sequence typing.
Fungal Genetics and Biology 45(4):400–421.
Byrnes, E. J., and J. Heitman. 2009. Cryptococcus gattii outbreak expands into the Northwestern
United States with fatal consequences. F1000 Biology Reports 1:62.
Byrnes, E. J., III, R. J. Bildfell, S. A. Frank, T. G. Mitchell, K. A. Marr, and J. Heitman. 2009. Mo-
lecular evidence that the range of the Vancouver Island outbreak of Cryptococcus gattii infection
has expanded into the Pacific Northwest in the United States. Journal of Infectious Diseases
199(7):1081–1086.
Byrnes, E. J., III, W. Li, Y. Lewit, H. Ma, K. Voelz, P. Ren, D. A. Carter, V. Chaturvedi, R. J. Bildfell,
R. C. May, and J. Heitman. 2010. Emergence and pathogenicity of highly virulent Cryptococcus
gattii genotypes in the northwest United States. PLoS Pathogens 6(4):e1000850.
Byrnes, E. J., W. Li, P. Ren, Y. Lewit, K. Voelz, et al. 2011. A diverse population of Cryptococcus
gattii molecular type VGIII in Southern California HIV/AIDS patients. PLoS Pathogens. ac-
cepted in principle pending revision.
Campbell, L. T., and D. A. Carter. 2006. Looking for sex in the fungal pathogens Cryptococcus neo-
formans and Cryptococcus gattii. FEMS Yeast Research 6(4):588–598.
Campbell, L. T., B. J. Currie, M. Krockenberger, R. Malik, W. Meyer, J. Heitman, and D. Carter.
2005a. Clonality and recombination in genetically differentiated subgroups of Cryptococcus
gattii. Eukaryotic Cell 4(8):1403–1409.
Campbell, L. T., J. A. Fraser, C. B. Nichols, F. S. Dietrich, D. Carter, and J. Heitman. 2005b. Clinical
and environmental isolates of Cryptococcus gattii from Australia that retain sexual fecundity.
Eukaryotic Cell 4(8):1410–1419.
244 FUNGAL DISEASES
Carter, D., N. Saul, L. Campbell, T. Bui, and M. Krockenberger. 2007. Sex in natural populations of
Cryptococcus gattii. In Sex in fungi: Molecular determination and evolutionary implications,
edited by J. Heitman, J. Kronstad, J. Taylor, and L. Casselton. Washington, DC: ASM Press.
Pp. 477–488.
Carter, D., L. Campbell, N. Saul, and M. Krockenberger. 2011. Sexual reproduction of Cryptococcus
gattii: A population genetics perspective. In Cryptococcus: From human pathogen to model
yeast, edited by J. Heitman, T. R. Kozel, J. K. Kwon-Chung, J. R. Perfect, and A. Casadevall.
Washington, DC: ASM Press. Pp. 299–311.
Chambers, C., L. MacDougall, M. Li, and E. Galanis. 2008. Tourism and specific risk areas for Cryp-
tococcus gattii, Vancouver Island, Canada. Emerging Infectious Diseases 14(11):1781–1783.
Chaturvedi, S., M. Dyavaiah, R. A. Larsen, and V. Chaturvedi. 2005. Cryptococcus gattii in AIDS
patients, southern California. Emerging Infectious Diseases 11(11):1686–1692.
Chen, S., T. Sorrell, G. Nimmo, B. Speed, B. Currie, D. Ellis, D. Marriott, T. Pfeiffer, D. Parr, and
K. Byth. 2000. Epidemiology and host- and variety-dependent characteristics of infection due
to Cryptococcus neoformans in Australia and New Zealand. Australasian Cryptococcal Study
Group. Clinical Infectious Diseases 31(2):499–508.
Cheng, P. Y., A. Sham, and J. W. Kronstad. 2009. Cryptococcus gattii isolates from the British Colum-
bia cryptococcosis outbreak induce less protective inflammation in a murine model of infection
than Cryptococcus neoformans. Infection and Immunity 77(10):4284–4294.
DeBess, E., et al. 2010. Emergence of Cryptococcus gattii—Pacific Northwest, 2004–2010. Morbidity
and Mortality Weekly Report 59:865–868.
Dromer, F., O. Ronin, and B. Dupont. 1992. Isolation of Cryptococcus neoformans var. gattii from an
Asian patient in France: Evidence for dormant infection in healthy subjects. Journal of Medical
and Veterinary Mycology 30(5):395–397.
D’Souza, C. A., J. W. Kronstad, G. Taylor, R. Warren, M. Yuen, G. Hu, W. H. Jung, A. Sham,
S. E. Kidd, K. Tangen, N. Lee, T. Zeilmaker, J. Sawkins, G. McVicker, S. Shah, S. Gnerre, A.
Griggs, Q. Zeng, K. Bartlett, W. Li, X. Wang, J. Heitman, J. E. Stajich, J. A. Fraser, W. Meyer,
D. Carter, J. Schein, M. Krzywinski, K. J. Kwon-Chung, A. Varma, J. Wang, R. Brunham, M.
Fyfe, B. F. Ouellette, A. Siddiqui, M. Marra, S. Jones, R. Holt, B. W. Birren, J. E. Galagan,
and C. A. Cuomo. 2011. Genome variation in Cryptococcus gattii, an emerging pathogen of
immunocompetent hosts. mBio 2(1):1–11.
Duncan, C., C. Stephen, S. Lester, and K. H. Bartlett. 2005. Sub-clinical infection and asymptomatic
carriage of Cryptococcus gattii in dogs and cats during an outbreak of cryptococcosis. Medical
Mycology 43(6):511–516.
Escandon, P., P. Ngamskulrungroj, W. Meyer, and E. Castaneda. 2007. In vitro mating of Colombian
isolates of the Cryptococcus neoformans species complex. Biomedica 27(2):308–314.
Fraser, J. A., R. L. Subaran, C. B. Nichols, and J. Heitman. 2003. Recapitulation of the sexual cycle
of the primary fungal pathogen Cryptococcus neoformans variety gattii: Implications for an
outbreak on Vancouver Island. Eukaryotic Cell 2:1036–1045.
Fraser, J. A., S. S. Giles, E. C. Wenink, S. G. Geunes-Boyer, J. R. Wright, S. Diezmann, A. Allen, J. E.
Stajich, F. S. Dietrich, J. R. Perfect, and J. Heitman. 2005. Same-sex mating and the origin of
the Vancouver Island Cryptococcus gattii outbreak. Nature 437(7063):1360–1364.
Georgi, A., M. Schneemann, K. Tintelnot, R. C. Calligaris-Maibach, S. Meyer, R. Weber, and P. P.
Bosshard. 2009. Cryptococcus gattii meningoencephalitis in an immunocompetent person 13
months after exposure. Infection 37(4):370–373.
Giles, S. S., T. R. Dagenais, M. R. Botts, N. P. Keller, and C. M. Hull. 2009. Elucidating the patho-
genesis of spores from the human fungal pathogen Cryptococcus neoformans. Infection and
Immunity 77:3491–3500.
Goldman, D. L., H. Khine, J. Abadi, D. J. Lindenberg, La Pirofski, R. Niang, and A. Casadevall.
2001. Serologic evidence for Cryptococcus neoformans infection in early childhood. Pediatrics
107(5):E66.
APPENDIX A 245
Hagen, F., S. van Assen, G. J. Luijckx, T. Boekhout, and G. A. Kampinga. 2010. Activated dormant
Cryptococcus gattii infection in a Dutch tourist who visited Vancouver Island (Canada): A
molecular epidemiological approach. Medical Mycology 48(3):528–531.
Heitman, J. 2006. Sexual reproduction and the evolution of microbial pathogens. Current Biology
16(17):R711–R725.
———. 2010. Evolution of eukaryotic microbial pathogens via covert sexual reproduction. Cell Host
and Microbe 8(1):86–99.
———. 2011. Cryptococcus from human pathogen to model yeast, edited by J. Heitman, T. R. Kozel,
J. K. Kwon-Chung, J. R. Perfect, and A. Casadevall. Washington, DC: ASM Press. P. 646.
Hoang, L. M., J. A. Maguire, P. Doyle, M. Fyfe, and D. L. Roscoe. 2004. Cryptococcus neoformans
infections at Vancouver Hospital and Health Sciences Centre (1997–2002): Epidemiology, mi-
crobiology and histopathology. Journal of Medical Microbiology 53(Pt 9):935–939e.
Hu, G., I. Liu, A. Sham, J. E. Stajich, F. S. Dietrich, and J. W. Kronstad. 2008. Comparative hybrid-
ization reveals extensive genome variations in the AIDS-associated pathogen Cryptococcus
neoformans. Genome Biology 9:R41.
Hull, C. M., and J. Heitman. 2002. Genetics of Cryptococcus neoformans. Annual Review of Genet-
ics 36:557–615.
Idnurm, A., Y. S. Bahn, K. Nielsen, X. Lin, J. A. Fraser, and J. Heitman. 2005. Deciphering the model
pathogenic fungus Cryptococcus neoformans. Nature Reviews Microbiology 3(10):753–764.
Kaufman, L., and S. Blumer. 1978. Cryptococcosis: The awakening giant. Proceedings of the Fourth
International Conference on the Mycoses: PAHO Scientific Publications No. 356. Pp. 176–184.
Kidd, S. E., F. Hagen, R. L. Tscharke, M. Huynh, K. H. Bartlett, M. Fyfe, L. Macdougall, T. Boekhout,
K. J. Kwon-Chung, and W. Meyer. 2004. A rare genotype of Cryptococcus gattii caused the
cryptococcosis outbreak on Vancouver Island (British Columbia, Canada). Proceedings of the
National Academy of Sciences, USA 101(49):17258–17263.
Kluger, E. K., H. K. Karaoglu, M. B. Krockenberger, P. K. Della Torre, W. Meyer, and R. Malik.
2006. Recrudescent cryptococcosis, caused by Cryptococcus gattii (molecular type VGII), over
a 13-year period in a Birman cat. Medical Mycology 44(6):561–566.
Kronstad, J. W., R. Attarian, B. Cadieux, J. Choi, C. A. D’Souza, E. J. Griffiths, J. M. Geddes, G.
Hu, W. H. Jung, M. Kretschmer, S. Saikia, and J. Wang. 2011. Expanding fungal pathogenesis:
Cryptococcus breaks out of the opportunistic box. Nature Reviews Microbiology 9(3):193–203.
Kwon-Chung, K. J. 1975. A new genus, Filobasidiella, the perfect state of Cryptococcus neoformans.
Mycologia 67:1197–1200.
———. 1976a. Morphogenesis of Filobasidiella neoformans, the sexual state of Cryptococcus neo-
formans. Mycologia 68 (4):821–833.
———. 1976b. A new species of Filobasidiella, the sexual state of Cryptococcus neoformans B and
C serotypes. Mycologia 68(4):943–946.
Kwon-Chung, K. J., J. C. Edman, and B. L. Wickes. 1992. Genetic association of mating types and
virulence in Cryptococcus neoformans. Infection and Immunity 60(2):602–605.
Lerner, C. W., and M. L. Tapper. 1984. Opportunistic infection complicating acquired immune defi-
ciency syndrome: Clinical features of 25 cases. Medicine (Baltimore) 63(3):155–164.
Lin, X., C. M. Hull, and J. Heitman. 2005. Sexual reproduction between partners of the same mating
type in Cryptococcus neoformans. Nature 434(7036):1017–1021.
Lindberg, J., F. Hagen, A. Laursen, J. Stenderup, and T. Boekhout. 2007. Cryptococcus gattii risk
for tourists visiting Vancouver Island, Canada. Emerging Infectious Diseases 13(1):178–179.
Littman, M. L., and L. E. Zimmer. 1956. Cryptococcosis. New York: Grune & Stratton, Inc.
Litvintseva, A. P., R. E. Marra, K. Nielsen, J. Heitman, R. Vilgalys, and T. G. Mitchell. 2003. Evi-
dence of sexual recombination among Cryptococcus neoformans serotype A isolates in sub-
Saharan Africa. Eukaryotic Cell 2(6):1162–1168.
Litvintseva, A. P., X. Lin, I. Templeton, J. Heitman, and T. Mitchell. 2007. Many globally isolated
AD hybrid strains of Cryptococcus neoformans originated in Africa. PLoS Pathogens 3(8):e114.
246 FUNGAL DISEASES
Upton, A., J. A. Fraser, S. E. Kidd, C. Bretz, K. H. Bartlett, J. Heitman, and K. A. Marr. 2007. First
contemporary case of human infection with Cryptococcus gattii in Puget Sound: Evidence for
spread of the Vancouver Island outbreak. Journal of Clinical Microbiology 45(9):3086–3088.
Velagapudi, R., Y. P. Hsueh, S. Geunes-Boyer, J. R. Wright, and J. Heitman. 2009. Spores as infectious
propagules of Cryptococcus neoformans. Infection and Immunity 77(10):4345–4355.
Vieira, J., E. Frank, T. J. Spira, and S. H. Landesman. 1983. Acquired immune deficiency in Haitians:
Opportunistic infections in previously healthy Haitian immigrants. New England Journal of
Medicine 308(3):125–129.
Wang, L., and X. Lin. 2011. Mechanisms of unisexual mating in Cryptococcus neoformans. Fungal
Genetics and Biology 48(7):651–660
West, S. K., E. J. Byrnes, S. Mostad, R. Thompson, R. Barnes, et al. 2008. Emergence of Crypto-
coccus gattii in the Pacific Northwest United States. Paper presented at the 48th meeting of
ICAAC/IDSA.
Xue, C., Y. Tada, X. Dong, and J. Heitman. 2007. The human fungal pathogen Cryptococcus can
complete its sexual cycle during a pathogenic association with plants. Cell Host and Microbe
1(4):263–273.
Xue, C., T. Liu, L. Chen, W. Li, I. Liu, J. W. Kronstad, A. Seyfang, and J. Heitman. 2010. Role of an
expanded inositol transporter repertoire in Cryptococcus neoformans sexual reproduction and
virulence. mBio 1(1):e00084–10.
Zimmer, B. L., H. O. Hempel, and N. L. Goodman. 1984. Pathogenicity of the basidiospores of
Filobasidiella neoformans. Mycopathologia 85(3):149–153.
A11
We lead inextricably mycotic lives: yeasts leaven our bread, ferment our
wine and beer, and inhabit our skins, mouths, and gastrointestinal tracts; however,
not all is harmony. Hippocrates reported aphthous ulcers consistent with thrush in
patients with severe debilitation, but it was not until the 1840s that in the newly
emerging field of clinical experimental medicine that thrush—as well as other
mycotic conditions, including ringworm—was recognized as being caused by
transmissible fungi. In 1923 Christine Marie Berkhout named what we now call
Candida albicans, for the white robe, toga candida, worn by Roman senators
and senatorial candidates (Emmons et al., 1977). Fungal infections in general,
and candida infections in particular, are important markers of innate or acquired
immune dysfunction. However, despite the estimated 1.5 million species of
fungi in the world, precious few cause human disease, and those that do (in the
38╛╛Originally published as Holland, Steven M; Vinh, Donald C.2009. Yeast Genetics on the Rise.
New England Journal of Medicine. Article DOI: 10.1056/NEJMe0907186. Copyright © 2009 Mas-
sachusetts Medical Society. Available at: http://www.nejm.org/doi/full/10.1056/NEJMe0907186.
39╛╛Steven M. Holland, M.D., and Donald C. Vinh, M.D. From the Laboratory of Clinical Infec-
tious Diseases, National Institute of Allergy and Infectious Diseases, National Institutes of Health,
Bethesda, MD.
APPENDIX A 249
Figure A11-1.eps
bitmap
APPENDIX A 251
FIGURE A11-1╇ (facing page). Mechanisms of Fungal Sensing and Control. The β-glucan
on the budding yeast forms of candida bind to dectin-1 on epithelial cells and phagocytes,
leading to activation of the CARD9 signaling complex (along with Bcl-10–MALT1)
(Panel A). This in turn produces cytokines that help drive CD4+ T lymphocytes toward the
Th17 phenotype, a STAT3-dependent process. Th17 lymphocytes elaborate interleukin-17
(augmenting neutrophil production and recruitment) and interleukin-22, which are syn-
ergic in the STAT3-dependent production of antimicrobial peptides by epithelial cells.
The generation of superoxide by the phagocytic NADPH oxidase system is crucial for
the killing of filamentous molds such as aspergillus (Panel B). Interactions between mac-
rophages and lymphocytes are required for the control of intracellular infections such as
the dimorphic fungi. In response to fungal infection, macrophages release interleukin-12,
which acts on its cognate receptor on T lymphocytes and natural killer (NK) lymphocytes.
Interleukin-12–stimulated lymphocytes release interferon-γ, which acts on macrophages
through STAT1 to kill intracellular fungi. Interferon-γ also augments tumor necrosis fac-
tor α (TNF-α), which acts in part through the nuclear factor κB (NF-κB) to kill fungi
and drive inflammatory responses. Molecules with identified mutations relevant to fungal
susceptibility are shown in red.
References
Beutler BA. TLRs and innate immunity. Blood 2009;113:1399–407.
Emmons CW, Binford CH, Utz JP, Kwon-Chung KJ. Medical mycology. 3rd ed. Philadelphia: Lea
& Febiger, 1977.
Ferwerda B, Ferwerda G, Plantinga TS, et al. Human dectin-1 deficiency and mucocutaneous fungal
infections. N Engl J Med 2009;361:1760–7.
Glocker E-O, Hennigs A, Nabavi M, et al. A homozygous CARD9 mutation in a family with suscep-
tibility to fungal infections. N Engl J Med 2009;361:1727–35.
Minegishi Y, Saito M, Nagasawa M, et al. Molecular explanation for the contradiction between
systemic Th17 defect and localized bacterial infection in hyper-IgE syndrome. J Exp Med
2009;206:1291–301.
A12
Infectious diseases of humans are receiving great attention due to their di-
rect and immediate impact on human mortality (King et al., 2006; WHO, 2008),
whereas the diseases of crop plants are subject to much less publicity although
they are threatening crop productivity, food security, and thereby the livelihood
of billions of people around the world. Currently, wheat rust fungi are among
the top 10 constraints for food production in many parts of the world, such as
sub-Saharan Africa, East Asia, and South Asia (FAO, 2010; Waddington et al.,
2010), mainly due to their epidemic potential and transboundary spread by wind
(Brown and Hovmøller, 2002).
According to U.S. Department of Agriculture statistics, global wheat produc-
tion was generally lower in 2010 compared to the previous year (USDA, 2011).
The reduced productivity was caused by several factors, such as unfavorable
weather conditions, resulting in large-scale flooding, droughts, and bushfires
in different parts of the world, as well as severe yellow rust epidemics (USDA,
2010).
Rust diseases have been recognized as being harmful to wheat since ancient
Greece. Theophrastus of Eressus (371–286 BC), the founder of botany and a
pupil of Aristoteles, noted in his book ΠΕΡΙ ΦΥΤΩΝ ΑΙΤΙΩΝ (The Causes of
Plants) that some plants of barley were more susceptible to the rusts than others,
the most susceptible denoted “Achilles Barley,” and he recognized the importance
of sowing in due time to reduce rust diseases on different crop plants such as
wheat, barley, peas, and beans (Theophrastus, 1990). Wheat rust may be caused
by three plant pathogenic fungi, Puccinia graminis (Leonard and Szabo, 2005),
Puccinia triticina (Bolton et al., 2008), and Puccinia striiformis (Hovmøller
et al., 2011). The corresponding diseases are termed black (stem) rust, brown
(leaf) rust, and yellow (stripe) rust, respectively (synonymous terms in brackets)
(Figure A12-1).
The three wheat rust fungi are biotrophic and heteroaecious, meaning they
generally require a primary wheat host for asexual reproduction and an alternate
host to complete sexual reproduction. Berberis spp. have been known to serve as
an alternate host for stem rust for more than two centuries. For instance, Schøler
(1818), who was a school teacher in Denmark, demonstrated a link between
stem rust on barberry and some cereals via repeated infection experiments. His
results started a so-called “barberry quarrel,” where the usefulness of barberry
eradication programs was discussed. Later, de Bary (1865, 1866) demonstrated
the complete life cycle of the stem rust fungus. However, the problems related to
barberry adjacent to cereal fields were observed by farmers in Europe much ear-
lier, with the first local laws against barberry being implemented in France in the
17th century and in the Americas in the 18th century (Roelfs, 1982). Surprisingly,
the discovery of barberry as an alternate host for yellow rust infecting wheat and
Poa grass, respectively, was not made until very recently (Jin et al., 2010). The
alternate host of P. triticina depend on the primary host: Isolates from cultivated
wheat and wild emmer have Thalictrum speciosissimum (in the Ranunculaceae)
as alternate host, whereas several species in the Boraginaceae, such as Anchusa
aggregata, Anchusa italica, Echium glomeratum, and Lycopsis arvensis, may
254 FUNGAL DISEASES
FIGURE A12-1╇ Typical macroscopic symptoms of rust infections on adult wheat plants.
Puccinia striiformis Westend (yellow/stripe rust) [left photo], Puccinia triticina (brown/
Figure
leaf rust) [center photo], and Puccinia A12-1.eps
graminis tritici (black/stem rust) [right photo].
bitmap
serve as alternate host for P. triticina from wild wheat and rye (Anikster et al.,
1997). Although this paper deals with all three wheat rusts, the primary focus is
yellow rust, which has spread at unprecedented scales in recent years and caused
severe epidemics even in areas where the disease was previously non-significant
or absent (Hovmøller et al., 2010).
rust exceeded 1 million metric tons over several years, and in each of the years
2003 and 2010 they mounted to 2.4 million tons (Long, 2000–2010), despite
increased and widespread use of agrochemicals already in 2003 (Chen, 2005). In
China, yellow rust is considered the most damaging disease on wheat, which is
grown on more than 20 million hectares (Wan et al., 2004). In three yellow rust
epidemic years, annual losses varied between 1.8 and 6 million tons; in 2002,
losses of 1.3 million tons were recorded and 1.9 million tons were saved by
treating at least 6 million hectares with fungicides (Wan et al., 2004). Yellow rust
epidemics reached record levels in northern Africa in 2009 (Ezzahiri et al., 2009)
and also in central and western Asia (Mboup et al., 2009), where more than 90
percent of important wheat varieties were susceptible to the disease (Sharma et
al., 2009). Areas particularly affected by yellow rust epidemics in 2009 and 2010
are illustrated in Figure A12-2.
New genetic variability of the pathogen, resulting from natural population
dynamic forces such as mutation, recombination, and migration, may be fol-
lowed by host-induced selection, which is a major driver of changes in gene and
genotypic frequencies of biotrophic plant pathogens (Hovmøller et al., 1997).
2009 2010
FIGURE A12-2╇ Map indicating the distribution of global wheat production and regions
of recent yellow rust epidemics.Figure
Each red A12-2.eps
dot is equivalent to 20,000 tons of wheat grain
production. The blue and orange circlesbitmap
indicate some of the areas in which severe yel-
low rust epidemics on wheat were reported in 2009 and 2010. Yellow rust severity varied
greatly from field to field within the highlighted areas, depending on susceptibility of
actual host cultivar, local environmental conditions, pathogen race distribution, and disease
management practices. The epidemics in Northwest Europe in 2009 (dotted circle) were
restricted to relatively few cultivars of wheat and triticale.
SOURCE: Adapted from Trethowan et al. (2002). Dr. Dave Hodson, Food and Agricul-
ture Organization of the United Nations, Rome, is acknowledged for kind permission to
reprint the map.
256 FUNGAL DISEASES
India (Prashar et al., 2007), Australia (Wellings, 2007; Wellings and McIntosh,
1990), and South Africa (Boshoff et al., 2002).
In addition to the virulence phenotype, pathogen aggressiveness, which is
a quantitative measure of the ability of a virulent isolate to cause disease on a
susceptible host plant, is a key determinant for the rate of pathogen spread and
evolution. Some of the recent epidemics since 2000 have been ascribed to the
emergence of yellow rust strains with increased aggressiveness and tolerance to
warm temperatures (Hovmøller et al., 2008). In this context, strain was defined
as a group of P. striiformis isolates sharing both molecular marker phenotype
and virulence phenotype. For instance, Milus et al. (2009) demonstrated a sig-
nificantly increased aggressiveness at both high and low temperatures, compared
with isolates of pre-2000 races from North America and Europe (Milus et al.,
2009). At the low-temperature regime, gradually changing from 10°C (night) to
18°C (day), the aggressive strains produced more than 70 percent more spores
per infected leaf area compared to reference isolates from before 2000. At the
high-temperature regime (12°C at night to 28°C during the day), isolates of the
aggressive strains produced approximately 150 percent more spores per infected
leaf area. Milus and colleagues (2009) concluded that the aggressive strain had
most likely enhanced the yellow rust epidemics in North America and contrib-
uted to the spread of yellow rust into areas that were previously considered too
warm for yellow rust epidemics. The aggressive strain detected in North America
was indistinguishable from a strain detected as exotic incursion in Australia in
2002 (Hovmøller et al., 2008; Wellings et al., 2003). In Europe, a nearly identi-
cal strain was first detected in 2000, whereas first appearance of this strain in
the Red Sea area and in western and central Asia is unknown (Hovmøller et al.,
2008). According to time and area, these observations may be comparable to the
famous panglobal spread of the Irish potato famine fungus Phytophthora infes-
tans (Goodwin et al., 1994).
The distinction between virulence and aggressiveness may be subject to con-
troversy because virulence toward a resistance gene with minor effects may be
manifest by increased disease progress and spore production. These are the same
variables used to measure aggressiveness. However, in case of virulence, there
should be evidence for a significant host–pathogen interaction, whereas in the
case of aggressiveness, the host–pathogen interaction is ideally non-significant.
This implies that for a range of host cultivars with varying levels of rust suscep-
tibility, they should rank the same for aggressive and non-aggressive isolates.
exotic incursions of yellow rust at continental scales since 1979, often resulting
in the spread of yellow rust epidemics to new regions or continents. Scherm and
Coakley (2003) noted that the rate of exotic pathogen invasion in the United
States had increased from about five instances per decade from 1940 to 1970
to more than three times that number during the 1990s. This increasing rate of
incursions of the yellow rust fungus may partly be ascribed to the emergence of
new, highly aggressive rust strains combined with increasing human travel and
commerce. However, regardless of the reasons for the increasing spread, the situ-
ation escalates, emphasizing the relevance of pathogen surveys covering larger
areas and a need for global coordination. The use of crop cultivars with similar
or identical resistance specificities across large wheat-growing areas supports
this conclusion.
The accelerating wheat yellow rust epidemics requires coordinated multina-
tional action to complement the ongoing initiatives of the Borlaug Global Rust
Initiative (www.globalrust.org) to fight the multivirulent strain of wheat stem
rust, known as Ug99 (Northoff, 2007, 2008). The speed by which threatening
new strains can be diagnosed and reported widely is essential, and the precision
by which traits such as virulence and aggressiveness are diagnosed is another
bottleneck (Hovmøller et al., 2011). Precise diagnosis of both traits require pure
and correct identification of the seed stocks of standard differential wheat variet-
ies used for the race assays, genetically pure pathogen isolates, and well-defined
experimental conditions for temperature, humidity, light, and plant nutrition, as
well as trained staff who can assess and interpret the virulence phenotype and
aggressiveness. The acknowledgment of these challenges led to the establish-
ment of a Global Rust Reference Center (GRRC) in 2008 targeting yellow rust,
supported by the International Center for Agricultural Research in the Dry Areas
(ICARDA), the International Maize and Wheat Improvement Center (CIMMYT),
and Aarhus University (Denmark) (Hovmøller et al., 2010). GRRC is accessible
for receiving yellow rust samples on a year-round basis, which is a major advance
complementing the capacities of existing regional and national rust diagnostic
laboratories. At the Borlaug Global Rust Initiative (BGRI) Coordination Meet-
ing in Syria in September 2009 (Dold, 2009), a proposal was made to extend the
activities of GRRC to all wheat rust fungi, that is, yellow rust, brown rust, and
stem rust.
The long-term aims of a Global Rust Reference Center would be the track-
ing of new wheat rust incursions and assessment of ongoing pathogen evolution
at global scales, supplying data to online early-warning systems for farmers,
breeders, and policy makers, like the current “Rust Spore” monitoring system
run by the Food and Agriculture Organization of the United Nations (FAO,
2010). Another main activity would be to assist in training of students and junior
scientists in rust pathology, and maintenance and extension of a unique world
wheat rust collection to facilitate resistance breeding efforts. The activities should
complement ongoing regional and national survey activities and research efforts
APPENDIX A 259
Concluding Remarks
The fight against infectious crop diseases has become a collective responsi-
bility and requires a collective investment with a global, long-term political and
scientific commitment. Such a global network will need to establish the physical
and human resources to support the progress of wheat rust management tech-
nologies, that is, pathogen monitoring and resistance breeding, training programs
for farmers and pathologists, and scientific progress. Hence a major priority of
a global network should be to link researchers and plant breeders (1) with each
other and with appropriate partners in resource-poor countries that are affected
most by sudden wheat rust epidemics, and (2) with political authorities to allow
for rapid actions. The establishment of the BGRI in 2006, inspired by Dr. Nor-
man Borlaug, the pioneer of the Green Revolution and the front leader of cereal
rust resistance breeding, represents a major milestone in this respect. The real
challenge is to ensure sustained activities beyond ongoing, short-term projects.
In the long term, surveillance and prevention measures of wheat rust epidemics
will only be successful if performed on a coordinated, global scale. Borlaug used
to say: “Rust never sleeps”—and events of recent years have shown how right
he was.
References
Bayles, R. A., K. Flath, M. S. Hovmøller, and C. de Vallavieille-Pope. 2000. Breakdown of the Yr17
resistance to yellow rust of wheat in northern Europe. Agronomie 20(7):805–811.
Bolton, M. D., J. A. Kolmer, and D. F. Garvin. 2008. Wheat leaf rust caused by Puccinia triticina.
Molecular Plant Pathology 9(5):563–575.
Boshoff, W. H. P., Z. A. Pretorius, and B. D. van Niekerk. 2002. Establishment, distribution, and
pathogenicity of Puccinia striiformis f. sp tritici in South Africa. Plant Disease 86(5):485–492.
Broad Institute. 2009. Genome sequencing of wheat stripe rust and comparative genomics of Puccinia
spp. U.S. Department of Agriculture. http://www.reeis.usda.gov/web/crisprojectpages/219263.
html. (accessed March 28, 2011).
Brown, J. K. M., and M. S. Hovmøller. 2002. Aerial dispersal of pathogens on the global and conti-
nental scales and its impact on plant disease. Science 297(5581):537–541.
Chen, W. Q., L. R. Wu, T. G. Liu, S. C. Xu, S. L. Jin, Y. L. Peng, and B. T. Wang. 2009. Race dynam-
ics, diversity, and virulence evolution in Puccinia striiformis f. sp tritici, the causal agent of
wheat stripe rust in China from 2003 to 2007. Plant Disease 93(11):1093–1101.
APPENDIX A 261
Chen, X. M. 2005. Epidemiology and control of stripe rust Puccinia striiformis f. sp tritici on wheat.
Canadian Journal of Plant Pathology-Revue Canadienne De Phytopathologie 27 (3):314–337.
Chen, X., and L. Penman. 2005. Stripe rust epidemic and races of Puccinia striiformis in the United
States in 2004. Phytopathology 95(6):S19.
de Bary, A. 1865. Neue Untersuchungen über die Uredineen, insbesondere die Entwicklung der
Puccinia. Monatsbericht der Koeniglich-Preussischen Akademie der Wissenschaften zu Berlin.
15–49.
———. 1866. Neue Untersuchungen über Uredineen, Zweite Mittheilung. Monatsberichten der
Akademie der Wissenschaften zu Berlin. 205–215.
De Vallavieille-Pope, C., H. Picardformery, S. Radulovic, and R. Johnson. 1990. Specific resistance
factors to yellow rust in seedlings of some French wheat varieties and races of Puccinia stri-
iformis Westend in France. Agronomie 10(2):103–113.
Dold, M. 2009. Top wheat experts call for scaling up efforts to combat Ug99 and other wheat rusts,
11 September 2009. http://www.eurekalert.org/pub_releases/2009-09/bc-twe091009.php (ac-
cessed March 28, 2011).
Enjalbert, J., X. Duan, M. Leconte, M. S. Hovmøller, and C. De Vallavieille-Pope. 2005. Genetic evi-
dence of local adaptation of wheat yellow rust (Puccinia striiformis f. sp tritici) within France.
Molecular Ecology 14(7):2065–2073.
Ezzahiri, B., A. Yahyaoui, and M. S. Hovmøller. 2009. An analysis of the 2009 epidemic of yellow
rust on wheat in Morocco. Paper presented at the Fourth Regional Yellow Rust Conference for
Central and West Asia and North Africa, Antalya, Turkey, October 11–12, 2009.
FAO (Food and Agriculture Organization of the United Nations). 2010. WHEAT RUST—Threat
to farmers and global food security. http://www.fao.org/agriculture/crops/core-themes/theme/
pests/wrdgp/en/ (accessed March 28, 2011).
Flath, K., and G. Bartels. 2002. Virulence situation in Austrian and German populations of wheat
yellow rust. Arbeitstagung 2001 der Vereinigung der Pflanzenzuchter und Saatgutkaufleute
Osterreichs gehalten vom 20. bis 22. November 2001 in Gumpenstein, Irdning, Austria. 51–56.
Verlag und Druck der Bundesanstalt ffn alpenllnstalt Landwirtschaft Gumpenstein.
Flor, H. H. 1956. The complementary genic systems in flax and flax rust. In Advances in Genetics,
edited by M. Demerec. New York: Academic Press.
Goodwin, S. B., B. A. Cohen, and W. E. Fry. 1994. Panglobal distribution of a single clonal lineage
of the Irish potato famine fungus. Proceedings of the National Academy of Sciences, USA
91(24):11591–11595.
Hovmøller, M. S. 2001. Disease severity and pathotype dynamics of Puccinia striiformis f. sp tritici
in Denmark. Plant Pathology 50(2):181–189.
Hovmøller, M. S., and K. E. Henriksen. 2008. Application of pathogen surveys, disease nurseries
and varietal resistance characteristics in an IPM approach for the control of wheat yellow rust.
European Journal of Plant Pathology 121(3):377–385.
Hovmøller, M. S., and A. F. Justesen. 2007a. Appearance of atypical Puccinia striiformis f. sp
tritici phenotypes in north-western Europe. Australian Journal of Agricultural Research 58
(6):518–524.
———. 2007b. Rates of evolution of avirulence phenotypes and DNA markers in a northwest Euro-
pean population of Puccinia striiformis f. sp tritici. Molecular Ecology 16:4637–4647.
Hovmøller, M. S., H. Ostergard, and L. Munk. 1997. Modelling virulence dynamics of airborne plant
pathogens in relation to selection by host resistance in agricultural crops. In The gene-for-gene
relationship in plant–parasite interactions, edited by I. R. Crute, E. B. Holub, and J. J. Burdan.
Oxfordshire: CAB International.
Hovmøller, M. S., A. F. Justesen, and J. K. M. Brown. 2002. Clonality and long-distance migration of
Puccinia striiformis f. sp tritici in North-West Europe. Plant Pathology 51(1):24–32.
Hovmøller, M. S., A. H. Yahyaoui, E. A. Milus, and A. F. Justesen. 2008. Rapid global spread of two
aggressive strains of a wheat rust fungus. Molecular Ecology 17(17):3818–3826.
262 FUNGAL DISEASES
Hovmøller, M. S., S. Walter, and A. F. Justesen. 2010. Escalating threat of wheat rusts. Science 329
(5990):369–369.
Hovmøller, M. S., C. K. Sørensen, S. Walter, and A. F. Justesen. 2011. Diversity of Puccinia
striiformis on cereals and grasses. Annual Review of Phytopathology 49. doi: 10.1146/
annurev-phyto-072910-095230.
Jin, Y., L. J. Szabo, and M. Carson. 2010. Century-old mystery of Puccinia striiformis life history
solved with the identification of Berberis as an alternate host. Phytopathology 100(5):432–435.
Johnson, R. 1992. Reflections of a plant pathologist on breeding for disease resistance, with emphasis
on yellow rust and eyespot of wheat. Plant Pathology 41(3):239–254.
Jørgensen, L. N., M. S. Hovmøller, J. G. Hansen, P. Lassen, B. Clark, R. Bayles, B. Rodemann,
M. Jahn, K. Flath, T. Goral, J. Czembor, P. du Cheyron, C. Maumene, C. de Pope, and G. C.
Nielsen. 2010. EuroWheat.org—A support to integrated disease management in wheat. Outlooks
on Pest Management 21(4):173–175.
King, D. A., C. Peckham, J. K. Waage, J. Brownlie, and M. E. J. Woolhouse. 2006. Infectious dis-
eases: Preparing for the future. Science 313(5792):1392–1393.
Kolmer, J. A. 2005. Tracking wheat rust on a continental scale. Current Opinion in Plant Biology
8(4):441–449.
Leonard, K. J., and L. J. Szabo. 2005. Stem rust of small grains and grasses caused by Puccinia
graminis. Molecular Plant Pathology 6(2):99–111.
Line, R. F., and A. Qayoum. 1992. Virulence, aggressiveness, evolution, and distribution of races
of Puccinia striiformis (the cause of stripe rust of wheat) in North America, 1968–87. USDA
Technical Bulletin 1788.
Long, D. L. 2000–2010. Small grain losses due to rust. http://www.ars.usda.gov/Main/docs.
htm?docid=10123 (accessed June 17, 2011).
Mboup, M., M. Leconte, A. Gautier, A. M. Wan, W. Chen, C. de Vallavieille-Pope, and J. Enjalbert.
2009. Evidence of genetic recombination in wheat yellow rust populations of a Chinese over-
summering area. Fungal Genetics and Biology 46(4):299–307.
McDonald, B. A., and C. Linde. 2002. Pathogen population genetics, evolutionary potential, and
durable resistance. Annual Review of Phytopathology 40:349–379.
Milus, E. A., K. Kristensen, and M. S. Hovmøller. 2009. Evidence for increased aggressiveness in
a recent widespread strain of Puccinia striiformis f. sp. tritici causing stripe rust of wheat.
Phytopathology 99(1):89–94.
Northoff, E. 2007. Wheat killer spreads from East Africa to Yemen: New partnership formed to
monitor and prevent spread of dangerous fungus. Food and Agriculture Organization of the
United Nations (FAO). http://www.fao.org/newsroom/en/news/2007/1000537/index.html (ac-
cessed March 28, 2011).
———. 2008. Wheat killer detected in Iran: Dangerous fungus on the move from East Africa to the
Middle East. Food and Agriculture Organization of the United Nations (FAO). http://www.fao.
org/newsroom/en/news/2008/1000805/index.html (accessed March 28, 2011).
Park, R., T. Fetch, D. Hodson, Y. Jin, K. Nazari, M. Prashar, and Z. Pretorius. 2010. International
surveillance of wheat rust pathogens—progress and challenges. http://www.globalrust.org/
db/attachments/about/19/1/BGRI%20oral%20papers%202010.pdf (accessed March 28, 2011).
Prashar, M., S. C. Bhardwaj, S. K. Jain, and D. Datta. 2007. Pathotypic evolution in Puccinia stri-
iformis in India during 1995–2004. Australian Journal of Agricultural Research 58(6):602–604.
Roelfs, A. P. 1982. Effects of barberry eradication on stem rust in the United States. Plant Disease
66(2):177–181.
Scherm, H., and S. M. Coakley. 2003. Plant pathogens in a changing world. Australasian Plant
Pathology 32(2):157–165.
Schøler, N. P. 1818. En afhandling om Berberissens skadelige Virkning på Sæden. Landoec. Tid. 8,
289–336 (in Danish).
APPENDIX A 263
Sharma R. C., A. Amanov, Z. Khalikulov, C. Martius, Z. Ziyaev, S. Alikulov. 2009. Wheat yellow
rust epidemic in Uzbekistan in 2009. Proceedings of the 4th Regional Yellow Rust Conference
for Central and West Asia and North Africa, Antalya, Turkey, October 10–12.
Stergiopoulos, I., and P. J. G. M. de Wit. 2009. Fungal effector proteins. Annual Review of Phytopa-
thology 47(1):233–263.
Theophrastus, E. 1990. De Causis Plantarum, books III IV, edited and translated by B. Einarson
and G. K. K. Link. Cambridge (Massachusetts), London (England): Harvard University Press.
Trethowan, R. M., D. Hodson, H.-J. Braun, W. H. Pfeiffer, and M. van Ginkel. 2002. Wheat breeding
environments. In Impacts of international wheat breeding research in the developing world,
1988–2002, edited by M. A. Lantican, H. J. Dubin, and M. L. Morris. Mexico, D.F.: CIMMYT.
USDA (U.S. Department of Agriculture). 2010. Middle East: Yellow rust epidemic affects regional
wheat crops. http://www.pecad.fas.usda.gov/highlights/2010/06/Middle%20East/ (accessed
March 28, 2011).
———. 2011. 2010/2011 wheat production. http://www.pecad.fas.usda.gov/ogamaps/default.cfm?
cmdty=Wheat&attribute=Production (accessed March 28, 2011).
Waddington, S., X. Li, J. Dixon, G. Hyman, and M. de Vicente. 2010. Getting the focus right: Produc-
tion constraints for six major food crops in Asian and African farming systems. Food Security
2(1):27–48.
Walter, S., E. Kemen, J. K. M. Brown, J. D. G. Jones, M. S. Hovmøller, and A. F. Justesen. 2009.
Omics approaches to understand the nature of virulence in Puccinia striiformis f.sp. tritici. Paper
presented at the 12th International Cereal Rusts and Powdery Mildews Conference, Antalya,
Turkey, October 13–16, 2009.
Wan, A., Z. Zhao, X. M. Chen, Z. He, S. Jin, Q. Jia, G. Yao, J. Yang, B. Wang, and G. Li. 2004.
Wheat stripe rust epidemic and virulence of Puccinia striiformis f. sp. tritici in China in 2002.
Plant Diseases 88:896–904.
Wan, A. M., X. M. Chen, and Z. H. He. 2007. Wheat stripe rust in China. Australian Journal of
Agricultural Research 58(6):605–619.
Wellings, C. R. 2007. Puccinia striiformis in Australia: A review of the incursion, evolution, and
adaptation of stripe rust in the period 1979–2006. Australian Journal of Agricultural Research
58(6):567–575.
Wellings, C. R., and R. A. McIntosh. 1990. Puccinia striiformis f.sp. tritici in Australasia: Pathogenic
changes during the first 10 years. Plant Pathology 39(2):316–325.
Wellings, C. R., D. G. Wright, F. Keiper, and R. Loughman. 2003. First detection of wheat stripe
rust in western Australia: Evidence for a foreign incursion. Australasian Plant Pathology
32(2):321–322.
WHO (World Health Organization). 2008. The World Health Report 2008—primary health care
(now more than ever), edited by T. Evans and W. Van Lerberghe. Geneva, Switzerland: WHO.
Zadoks, J. C. 1961. Yellow rust on wheat studies in epidemiology and physiologic specialization.
European Journal of Plant Pathology 67(3):69–256.
264 FUNGAL DISEASES
A13
Introduction
Although the vast majority of fungal species do not cause disease, ones that
affect plants are responsible for significant economic losses (Skamnioti and Gurr,
2009). In the past couple of decades, fungal diseases of humans have become an
increasing threat, especially for people who are immunologically compromised
(Sexton and Howlett, 2006). Consequently there has been a rapid explosion in
knowledge about fungal pathogenesis of animals and this has been taken up by
people studying plant pathogenic fungi. Pathogenesis involves the interaction of
two partners with input from the environment, a concept described as the “dis-
ease triangle” in plant pathology. The “damage-response” concept developed for
animal pathogens emphasizes that the outcome of an interaction is determined
by the amount of damage incurred on the host (Casadevall and Pirofski, 2003).
These concepts are useful reminders of the complexity of the interaction and the
interdependence of host and pathogen.
Fungi enter plants via the stomatal apertures, where air exchange occurs;
by digesting the cuticle and cell wall with hydrolytic enzymes; or by develop-
ing infection structures, appressoria, which accumulate high concentrations of
glycerol and puncture the surface due to high turgor pressure (Van de Wouw and
Howlett, 2011). As mentioned above, fungi enter animals through the skin or via
inhalation or wounds; an exception is Histoplasma capsulatus, which enters the
host via receptor-mediated endocytosis (Woods, 2003).
The first basal or innate layer of defense is very similar in plants and ani-
mals. It involves binding of pathogen associated molecular patterns (PAMPs) to
pattern recognition receptors on the host membrane surface. This triggers signal-
ing pathways that induce a range of defense responses, including production of
reactive oxygen species and sometimes programmed cell death. This response
is termed “pathogen triggered immunity.” PAMPs common to plant and animal
pathogenic fungi include wall carbohydrate fragments, such as chitin oligosac-
charides or b 1,3 glucans. As well as conservation among PAMPs, there is also
conservation in structural domains of Pattern Recognition Receptors (for review
see Zipfel, 2009).
In plants often the responses triggered by innate basal immunity are not
strong enough to stop pathogen invasion. Consequently a second round of rec-
ognition occurs involving effector-triggered immunity (Jones and Dangl, 2006).
Often there is a “gene for gene” interaction between avirulence (effector) genes
in the pathogen and resistance genes in the plant, such that the pathogen is unable
to attack host genotypes with the corresponding resistance genes. Thus there is
usually a high degree of host specificity with plant diseases, with only particular
varieties (genotypes) of a single species being susceptible. In contrast to plant
diseases, most animal diseases do not display host genotype specificity, although
some fungi only cause disease in certain animal species. Apart from innate basal
immunity, the immune system of mammals is very different from that of plants.
The responses that come into play if the pathogen is not stopped by the innate
basal immunity response include the innate complement system, circulating cells
such as phagocytes that can internalize and destroy pathogen cells, and adaptive
antibody-mediated defenses (Speth et al., 2004).
Fungal plant pathogens have a range of lifestyles and nutritional require-
ments. Many have a saprophytic phase, and some are obligate, surviving only on
their hosts. Some (biotrophs) require hosts to be living, while others (necrotrophs)
kill plant tissues. Some fungi are both biotrophs and necrotrophs at different
stages during growth in planta. Some fungi such as mycorrhizae and endophytes
colonize plants in a symbiotic relationship deriving carbon from photosynthesis
by the plant and often conferring drought tolerance on the plant. Fungi that infect
animals often have a saprophytic lifestage in the soil, and few if any are obligate.
After invasion, fungi then need to colonize the host, derive nutrition, and
avoid or subvert host defense responses. In many plant–pathogen interactions, de-
fense responses include the hypersensitive response, whereby an oxidative burst
268 FUNGAL DISEASES
by the plant generates reactive oxygen species associated with the programmed
cell death of host cells. This can arrest pathogen growth, particularly that of bio-
trophs. Necrotrophic fungi can subvert this defense process to derive nutrition
from the dead host tissue. Toxins produced by necrotrophic fungi also kill plant
tissue. However, they are generally not often important disease determinants.
This is also often the case for secondary metabolite toxins produced by fungi
that attack animals. Their role may be to protect fungi against predators such as
insects, nematodes, and amoebae during their saprophytic growth phase in the
soil (Kempken and Rohlfs, 2010).
To complete its lifecycle, a fungus must reproduce and exit the host to find
another host. Many animal pathogenic ascomycetes do not have a sexual phase.
However, the sexual cycle of basidiomycetes such as Cryptococcus spp. is ex-
tremely important in virulence, as discussed by Heitman. 42 Both animal- and
plant-pathogenic fungi reproduce mitotically within the host. Although plant-to-
plant transmission of fungal disease is very common, direct transmission of fun-
gal pathogens between mammalian hosts is unusual. In plant pathogens, conidia
(vegetative spore) production usually occurs after infection has been established
and lesions have developed. Conidia are spread from plant to plant by wind or
in water droplets. However, many pathogens of mammals are transmitted by
inhalation as hyphae or as arthroconidia, in dust and wind (e.g., C. immitis) or
soil. Increasingly, biofilms, whereby a community of microorganisms attaches
to a solid surface, mediate dispersal of fungi in hospital environments (Ramage
et al., 2009).
Closing Remarks
The development of fungal diseases of plants and animals has many parallels.
A fungus must overcome many hurdles to successfully invade a plant and cause
APPENDIX A 271
References
Bartlett, D. W., J. M. Clough, J. R. Godwin, A. A. Hall, M. Hamer, and B. Parr-Dobrzanski. 2002.
The strobilurin fungicides. Pest Management Science 58:649–652.
Beckers, G. J., and U. Conrath. 2007. Priming for stress resistance: From the lab to the field. Current
Opinion in Plant Biology 10:425–431.
Bergman, A., and A. Casadevall. 2010. Mammalian endothermy optimally restricts fungi and meta-
bolic costs. MBIO 1(5):e00212–10.
Brefort, T., G. Doehlemann, A. Mendoza-Mendoza, S. Reissmann, A. Djamei, and R. Kahmann. 2009.
Ustilago maydis as a pathogen. Annual Review of Phytopathology 47:423–445.
Cairns, T., F. Minuzzi, and E. Bignell. 2010. The host-infecting fungal transcriptome. FEMS Micro-
biology Letters 307:1–11.
Carpita, N. C., and D. M. Gibeaut 1993. Structural models of primary-cell walls in flowering plants—
consistency of molecular structures with the physical properties of the wall during growth. Plant
Journal 3:1–30.
Casadevall, A., and L.A. Pirofski. 2003. The damage–response framework of microbial pathogenesis.
Nature Review Microbiology 1:17–24.
Ebbole, D. J. 2007. Magnaporthe as a model for understanding host–pathogen interactions. Annual
Review of Phytopathology 45:437–456.
Fudal, I., S. Ross, L. Gout, F. Blaise, M. L. Kuhn, M. R. Eckert, L. Cattolico, S. Bernard-Samain,
M. H. Balesdent, and T. Rouxel. 2007. Heterochromatin-like regions as ecological niches for
avirulence genes in the Leptosphaeria maculans genome: Map-based cloning of AvrLm6. Mo-
lecular Plant Microbe Interactions 20:459–470.
Gout, L., I. Fudal, M. L. Kuhn, F. Blaise, M. Eckert, L. Cattolico, M.-H. Balesdent, and T. Rouxel.
2006. Lost in the middle of nowhere: The AvrLm1 avirulence gene of the Dothideomycete
Leptosphaeria maculans. Molecular Microbiology 60:67–80.
Jones, J. D. G., and J. Dangl. 2006. The plant immune system. Nature 444:323–329.
Kempken, F., and M. Rohlfs. 2010. Fungal secondary metabolite biosynthesis—a chemical defence
strategy against antagonistic animals? Fungal Ecology 3:107–114.
Martínez-Rocha, A. L., M. I. G. Roncero, A. López-Ramirez, M. Mariné, J. Guarro, G. Martínez-
Cadena, and A. Di Pietro. 2008. Rho1 has distinct functions in morphogenesis, cell wall biosyn-
thesis and virulence of Fusarium oxysporum. Cellular Microbiology 10:1339–1351.
McDonald, B. A., and C. Linde. 2002. Pathogen population genetics, evolutionary potential, and
durable resistance. Annual Review of Phytopathology 40:349–379.
Mylonakis, E., A. Casadevall, and F. M. Ausubel. 2007. Exploiting amoeboid and non-vertebrate
animal model systems to study the virulence of human pathogenic fungi. PLoS Pathogens
3:859–865.
272 FUNGAL DISEASES
Woods, J. 2003. Knocking on the right door and making a comfortable home: Histoplasma capsula-
tum intracellular pathogenesis. Current Opinion in Microbiology 6:327–331.
Zipfel, C. 2009. Early molecular events in PAMP-triggered immunity. Current Opinion in Plant
Biology 12:414–420.
A14
Climate change is likely to become a major issue in plant pathology over the
coming years. In this overview, we provide recent evidence for this statement.
Moreover, we point out the importance for future plant disease management of
climate change interactions with other global change drivers, such as increased
long-distance trade. Recent advances in aerobiology, together with new molecu-
lar tools used in landscape and geographical genetics, can help in addressing the
challenges posed by an increasing number of emerging plant diseases worldwide.
There is increasing evidence that climate change will be a key issue in how plant
pathogens will affect food security and ecosystem health.
Less knowledge is available on the potential impacts of climate change on
biological control of exotic fungal plant pathogens. Network theory is a promis-
ing tool to improve biosecurity in the face of the increased volumes of traded
plants coupled with climate warming. Although there are now many reviews of
the literature on the topic of climate change and plant diseases, there is a need to
keep up with the rapid development of the subject.
Introduction
The emergence of fungal plant pathogens can occur as the result of at least
three processes: (1) the evolution of new genotypes within a pathogen population
that is endemic to a habitat or location (e.g., through genetic recombination); (2)
the introduction of an exotic pathogen genotype into a new, receptive habitat or
location; and/or (3) the natural selection of new pathogen genotypes from an
endemic population as a result of changes in host genotype, host species, or ex-
ternal pressures (e.g., environmental conditions, fungicide applications, changes
in plant cultural practices). Emergence is a function of introduction (via dispersal
or evolution), establishment (i.e., adaptation to the habitat or location), and spread
44╛╛Division
of Biology, Imperial College London, Silwood Park, Ascot, SL5 7PY, U.K.
45╛╛Department
of Plant Pathology, Kansas State University, 4024 Throckmorton Plant Sciences
Center, Manhattan, KS 66506���-5502.
274 FUNGAL DISEASES
FIGURE A14-1╇ The increase in goods (109 tons × km) moved in the United Kingdom
from the 1930s to the 1990s. Figure A14-1.eps
SOURCE: Schulz (2004).
bitmap
plant diseases such as wheat stripe rust, wheat stem rust, potato late blight, and
Southern corn leaf blight. Much of the earlier literature on this approach is cited
by the authors. They found that the estimated power law parameter varied little
over five orders of magnitude on a distance scale. Evidence was found to support
the hypothesis that disease advances through accelerating, rather than constant,
waves. Integration of (unmanned) aerial measurements of spore concentration at
various distances from the source with simulation of spore flight trajectories is
important to develop reliable decision support systems to predict risk of disease
spread, as shown for Phytophthora infestans in potato fields in Virginia (Aylor
et al., 2011; Techy et al., 2010).
A feature of recent work on dispersal has been the integration with popula-
tion genetics aspects of pathogen diversity (Hovmøller et al., 2008; Montarry
et al., 2010). This can operate at the level of host specialization and biotrophy and
the ways in which wind dispersal acts as a survival mechanism—what has been
termed “Oases in the desert” (Brown et al., 2002). Equally, the strongly stochastic
nature of long-distance dispersal can lead to founder effects in the pathogen popu-
lation (Brown and Hovmøller, 2002). In such cases, pathogen genotypes that suc-
cessfully establish in new regions and/or on new cultivars may be different from
those at the source—the so-called founder effect. Such effects were found in the
migration of the fungus causing black Sigatoka of bananas, Mycosphaerella fi-
jiensis, from Southeast Asia to Africa and Latin America (Brown and Hovmøller,
2002; Rivas et al., 2004). In both of the new regions founder effects were present
in the new invasions. Unresolved for this pathogen is the relative importance of
air-dispersed ascospores (probably limited) and the movement of infected plant
material (largely unrecorded). The fungus Corynospora cassiicola has a wide
geographical range in the tropics and subtropics and many plant hosts. Common
fungal lineages were widely distributed geographically, indicating long-distance
dispersal of clonal lineages, but also previously unrecognized genetic diversity
involving some degree of host specialization on some hosts (Dixon et al., 2009).
The advent of modern molecular tools in epidemiology provides a step change
in both the tracking of dispersal of novel fungal genotypes and in risk assess-
ment of emerging fungal diseases in plants and in animals (Gladieux et al., 2011;
Moslonka-Lefebvre et al., 2011).
troma needle blight epidemic following the devastating mountain pine beetle
outbreaks. The resilience of northern boreal forests to rapid climate change can
be questioned (Chapin et al., 2010), as can forest establishment under condi-
tions of permafrost thaw where fungal pathogens may affect seedling survival
(Camill et al., 2010). In forest tree nurseries in Finland, rust, powdery mildew,
and other fungal leaf diseases are already causing more problems because of
climate warming (Lilja et al., 2010). A more optimistic outlook due to improved
adaptive management practices is presented with respect to white pine blister
rust (Hunt et al., 2010), where white pines have broad ecological ranges and are
less likely to be maladapted thus succumbing to the disease, and hence may be
more resilient in the long term. Similar issues will need to be faced with regard
to urban trees (Tubby and Webber, 2010) and Mediterranean forests (Attorre
et al., 2011; La Porta et al., 2008), where the impact of non-native insect pests
and fungal pathogens introduced through trade pathways (see later sections) is
already being observed.
The above discussion concerns general issues regarding climate change and
plant diseases. There have been many accounts of global/climate change impacts
on diseases caused by specific fungal plant pathogens in recent years. A selection
of these is cited with summary comments in Table A14-1.
TABLE A14-1╇ Selected Papers Illustrating the Effects of Climate and Global
Change Factors on Specific Pathogen–Host Systems
Pathogen Host Reference Comments
Valsa Alder Worrall Previous oscillation period in damage of ~21 years
melandiscus et al. likely to dampen with warmer summers and with
(2010) no period of recovery
Erisiphe Grapevine Pugliese Increase in CO2 concentration did not affect
necatrix et al. incidence possibly due to increased photosynthesis
(2010) with higher CO2 and temperature
Phytophthora Beech Fleishmann Elevated CO2 and low N supply enhanced
citricola et al. susceptibility, but host compensation followed
(2010)
Cercospora Red bud and McElrone Incidence/severity under elevated CO2 greater
spp. Sweet bud et al. with above average rainfall and temperature (one
(2010) species), but mitigated by higher photosynthetic
efficiency
Leptosphaeria Oilseed rape Stonard Geographic variation in species may be exacerbated
maculans and et al. by climate change with more damaging species
L. biglobosa (2010) expanding in range
Didymella Chickpea Frenkel Isolates infecting crop and wild Cicer show
rabiei et al. different adaptation to high temperature, with the
(2010) potential for hybrids infecting both hosts over a
broader temperature range
Phakopsora Soybean Del Ponte Integrating epidemiological and meteorological
pachyrhizi and Esker factors suggest no restricted overwintering areas in
(2008) Brazil
Seiridium Mediterranean Zocca et al. Pathogen is associated with insect vectors, which
cardinale cypress (2008) are able to reach the range margin and thus the
continuous threat of arrival in the expanding range
Cronartium White pine Frank et al. Pathogen arrived in New Mexico in ~1970 with
ribicola (2008) upper flow synoptic classification indicating early
June 1969 most favourable
Fusarium Wheat Xu et al. Environmental conditions differentially affect
spp. (2008) infection and colonization process and the
comparative abundance of six toxigenic species in
the head blight disease
Fusarium Maize Horvath Damage by beetles provides conducive growing
spp. (2003) conditions for the toxigenic fungi, with feeding
behavior changing under drought conditions
APPENDIX A 281
FIGURE A14-2╇ The world in 1897, with British possessions marked in red.
Figure A14-2.eps
SOURCE: Cambridge University Library available at Wikimedia Commons.
bitmap
et al., 2010; Perrings et al., 2010). Emerging plant diseases can be seen as nega-
tive externalities deriving from the international trade in plants (Lansink, 2011).
The removal of trade barriers, however, can be beneficial and in some cases
desirable. For example, seed trade legislation was designed primarily to protect
trade and return royalties to contemporary plant breeders. Increasingly, the im-
portance of exploiting the genetic diversity present in cereal land races has been
recognized, but to exploit their use fully, changes in legislation will be required
(Newton et al., 2010). Also the genetic diversity of target plant pathogens should
be used in building comprehensive collections to allow efficient, reliable, and
specific diagnostic and detection tools in the national and international trade
(Barba et al., 2010).
The Sanitary and Phytosanitary Agreement of the World Trade Organisation
specifies that countries cannot regulate against unknown pests, yet many invasive
alien forest insects and pathogens were new to science when first recorded in a
new environment (Britton et al., 2010). To counter this, effective surveillance
systems are required to facilitate early detection; these are lacking in many na-
tions. Britton et al. (2010) propose a global network of sentinel plantings based
in historical gardens and arboreta to enable early detection and rapid response
APPENDIX A 283
Dispersal Associated with the Introduction of Novel Crops and Plant Species
Farmers world-wide tend to maximize the potential of the crops they grow,
in ways that are appropriate to their socioeconomic circumstances, but often
only sensible in the short term or not considering side effects such as enhanced
opportunities for plant pathogens. For example, modern agriculture relies heav-
ily on increasing the size of fields to enable a more efficient mechanization and
economies of scale. But this leads to more uniform landscapes, with potential
repercussions on epidemic thresholds of plant pathogens (Moslonka-Lefebvre
et al., 2011). Not just in natural ecosystems (Scherrer and Körner, 2011), but
probably also in agriculture, small-scale topographic variability will be important
for plant species to be able to cope with climate change. In some cases, there
may be opportunities to grow a novel crop because the changing climate makes
production possible in areas not previously suitable. This may lead to hosts for
plant pathogens that were previously unavailable.
In other cases, the cultivation of novel crops arises from other, often politi-
cal inducements: for example, the requirement that each country in the Euro-
pean Union must derive a certain percentage of its energy use from renewable
sources. Hence, there has been a diversion of food crop (cereal) production to
bioenergy use, and the planting of non-food crops such as willow, and other
short-rotation coppiced trees, and grasses such as Miscanthus for such use. As
pointed out by Stewart and Cromey (2010), new disease threats, often from
fungal plant pathogens, are likely to emerge among the existing ones. This is
likely to occur because the crop is new or because the cultivation method (e.g.,
high-density monoculture) is different from previous practice. In conservation
biology, there is a hefty discussion on the issue of assisted migration (managed
relocation), which is thought to be going to be necessary for the many species
not able to track the predicted rapid variations in climate, although it will result
in artificial shifts in species distributions and ecosystems (Loss et al., 2011).
In such debates, little attention has been paid to the likelihood that assisted
migration will result in unwanted introductions of, for example, exotic plant
pathogens.
286 FUNGAL DISEASES
Concluding Comments
In cropland, forests, and grasslands, experimentation is needed at the land-
scape level to investigate plant health adaptation approaches to climate change
(Holdenrieder et al., 2004). Similar experimentation, if accompanied by subse-
quent data analysis and dissemination, is also likely to be beneficial for pathway
288 FUNGAL DISEASES
control and diagnostic systems (Albers et al., 2010; Norton and Taylor, 2010).
Monitoring of plant pathosystems and associated organisms under changing
conditions is key to refine models, so as to successfully predict further changes
(Luedeling et al., 2011). Similarly, review of the burgeoning literature on the
subjects of climate change, land use, and plant health is essential and should be
funded on a long-term basis (Jeger and Pautasso, 2008; Pautasso et al., 2010a).
Analyses of past and current trends and simulations of future plant health sce-
narios are likely to be more influential and effective if accompanied by stake-
holder involvement and interactions with policy experts (Dwyer, 2011; Mills
et al., 2011). Climate change and other global change drivers are not the only
factors involved in plant disease epidemiology, so it is still important to consider
the important role of, for example, improvements in agronomic and silvicultural
practices in reducing current disease potential, whether or not the climate is
likely to become more conducive to a certain chronology of diseases in the com-
ing decades (Savary et al., 2011). Nonetheless, it is time to start adapting to the
challenges to plant health posed by the future climate change scenarios (Juroszek
and von Tiedemann, 2011; Legrève and Duveiller, 2010; Olesen et al., 2011;
Planinšek et al., 2011).
The emergence of fungal plant pathogens, facilitated by climate change
and globalization, will challenge our ability to meet future food demands for
a growing population and to achieve sustainable environments that provide the
ecosystem services on which societies depend. Policies informed by science and
an infrastructure that supports plant health are prerequisites to that future.
Acknowledgments
Many thanks to C. Brasier, B. D. L. Fitt, M. Garbelotto, K. Garrett, O.
Holdenrieder, M. W. Shaw, and J. Webber for insights and discussions. This
overview was partly funded by the Rural Economy and Land Use Programme
(RELU), U.K.
References
Agosta, S. J., N. Janz, and D. R. Brooks. 2010. How specialists can be generalists: Resolving the
“parasite” paradox and implications for emerging infectious disease. Zoologia 27:151–162.
Albers, H. J., C. Fischer, and J. N. Sanchirico. 2010. Invasive species management in a spatially
heterogeneous world: Effects of uniform policies. Resource and Energy Economics 32:483–499.
Albrectsen, B. R., J. Witzell, K. M. Robinson, S. Wulff, V. M. C. Luquez, R. Ågren, and S. Jansson.
2010. Large scale geographic clines of parasite damage to Populus tremula. L. Ecography
33:483–493.
Alexander, H. M. 2010. Disease in natural plant populations, communities, and ecosystems: Insights
into ecological and evolutionary processes. Plant Disease 94:492–503.
Andrade, D., Z. T. Pan, W. Dannevik, and J. Zidek. 2009. Modeling soybean rust spore escape from
infected canopies: Model description and preliminary results. Journal of Applied Meteorology
and Climatology 48:789–803.
APPENDIX A 289
Attorre, F., M. Alfò, M. De Sanctis, F. Francescani, R. Valenti, M. Vitale, and F. Bruno. 2011. Evalu-
ating the effects of climate change on tree species abundance and distribution in the Italian
peninsula. Applied Vegetation Science 14:242–255.
Aukema, J. E., D. G. McCullough, B. Von Holle, A. M. Liebhold, K. Britton, and S. J. Frankel. 2010.
Historical accumulation of nonindigenous forest pests in the continental United States. Biosci-
ence 60:886–897.
Awad, A. H. A. 2005. Vegetation: A source of air fungal bio-contaminant. Aerobiologia 21:53–61.
Aylor, D. E., D. G. Schmale, E. J. Shields, M. Newcomb, and C. J. Nappo. 2011. Tracking the potato
late blight pathogen in the atmosphere using unmanned aerial vehicles and Lagrangian model-
ing. Agriculture and Forest Meteorology 151:251–260.
Baeten, L., P. De Frenne, K. Verheyen, B. J. Graae, and M. Henry. 2010. Forest herbs in the face of
global change: A single-species-multiple-threats approach for Anemone nemorosa. Plant Ecol-
ogy and Evolution 143:19–30.
Barba, M., I. Van den Bergh, A. Belisario, and F. Beed. 2010. The need for culture collections to sup-
port plant pathogen diagnostic networks. Research in Microbiology 161:472–479.
Beggs, P. J. 2010. Adaptation to impacts of climate change on aeroallergens and allergic respiratory
diseases. International Journal of Environmental Research and Public Health 7:3006–3021.
Berger, C. N., S. V. Sodha, R. K. Shaw, P. M. Griffin, D. Pink, P. Hand, and G. Frankel. 2010. Fresh
fruit and vegetables as vehicles for the transmission of human pathogens. Environmental Mi-
crobiology 12:2385–2397.
Boa, E. 2007. Plant healthcare for poor farmers: An introduction to the work of the Global Plant
Clinic. APSnet http://www.apsnet.org/publications/apsnetfeatures/Pages/PoorFarmers.aspx (ac-
cessed June 13, 2011).
Brasier, C. M. 2008. The biosecurity threat to the U.K. and global environment from international
trade in plants. Plant Pathology 57:792–808.
Brasier, C., and J. Webber. 2010. Sudden larch death. Nature 466:824–825.
Britton, K. O., P. White, A. Kramer, and G. Hudler. 2010. A new approach to stopping the spread
of invasive insects and pathogens: Early detection and rapid response via a global network of
sentinel plantings. New Zealand Journal of Forestry Science 40:109–114.
Brown, J. K. M., and M. S. Hovmøller. 2002. Aerial dispersal of pathogens at the global and conti-
nental scale and its impact on plant disease. Science 297:537–541.
Brown, J. K. M., M. S. Hovmøller, R. A. Wyand, and D.-Z. Yu. 2002. Oases in the desert: Dispersal
and host specialization of biotrophic fungal pathogens of plants. In Dispersal, edited by J. M.
Bullock, R. E. Kenwood, and R. S. Hails. Oxford, England: Blackwell Science. Pp. 395–409.
Busby, P. E., and C. D. Canham. 2011. An exotic insect and pathogen disease complex reduces
aboveground tree biomass in temperate forests of eastern North America. Canadian Journal of
Forest Research 41:401–411.
Butterworth, M. H., M. A. Semenov, A. Barnes, D. Moran, J. S. West, and B. D. L. Fitt. 2010. North–
South divide: Contrasting impacts of climate change on crop yields in Scotland and England.
Journal of the Royal Society Interface 7:123–130.
Camill, P., L. Chihara, B. Adams, C. Andreass, A. Barry, S. Kalim, J. Limmer, M. Mandell, and G.
Rafert. 2010. Early life history transitions and recruitment of Picea mariana in thawed boreal
permafrost peatlands. Ecology 91:448–459.
Chakraborty, S., and A. C. Newton. 2011. Climate change, plant diseases and food security: An
overview. Plant Pathology 60:2–14.
Chapin, F. S., A. D. McGuire, R. W. Ruess, T. N. Hollingsworth, M. C. Mack, J. F. Johnstone, E. S.
Kasische, E. S. Euskirchen, J. B. Jones, M. T. Jorensen, K. Kielland, G. P. Kofinas, M. R.
Turetsky, J. Yarie, A. H. Lloyd, and D. L. Taylor. 2010. Resilience of Alaska’s boreal forest to
climatic change. Canadian Journal of Forest Research 40:1360–1370.
Compant, S., M. G. A. van der Heijden, and A. Sessitsch. 2010. Climate change effects in beneficial
plant–microorganism interactions. FEMS Microbiology Ecology 73:197–214.
290 FUNGAL DISEASES
Conway, D., and E. L. F. Schipper. 2011. Adaptation to climate change in Africa: Challenges and
opportunities identified. Global Environmental Change 21:227–237.
Crooks, J. A. 2005. Lag times and exotic species: The ecology and management of biological inva-
sions in slow-motion. Ecoscience 12:316–329.
Dallafior, T. N., and A. Sesartic. 2010. Global fungal spore emissions, review and synthesis of litera-
ture data. Biogeosciences Discussions 7:8445–8475.
D’Amato, G., and L. Cecchi. 2008. Effects of climate change on environmental factors in respiratory
allergic diseases. Clinical & Experimental Allergy 38:1264–1274.
Davis, F. W., M. Borchert, R. K. Meentenmeyer, A. Flint, and D. M. Rizzo. 2010. Pre-impact forest
composition and ongoing tree mortality associated with sudden oak death in the Big Sur region,
California. Forest Ecology and Management 259:2342–2354.
Dehnen-Schmutz, K., O. Holdenrieder, M. J. Jeger, and M. Pautasso. 2010. Structural change in the
international horticultural industry: Some implications for plant health. Scientia Horticulturae
125:1–15.
Del Ponte, E. M., and P. D. Esker. 2008. Meteorological factors and Asian soybean rust epidemics—a
systems approach and implications for risk assessment. Scientia Agricola 65:88–97.
Deliopoulos, T., P. S. Kettlewell, and M. C. Hare. 2010. Fungal disease suppression by inorganic salts:
A review. Crop Protection 29:1059–1075.
Desprez-Loustau, M.-L., R. Courtecuisse, C. Robin, C. Husson, P.-A. Moreau, D. Blancard, M.-A.
Selosse, B. Lung-Escarmant, D. Piou, and I. Sache. 2010. Species diversity and drivers of
spread of alien fungi (sensu lato) in Europe with a particular emphasis on France. Biological
Invasions 12:157–172.
Dixon, B. 2005. Trees become casualties of war. The Lancet 5:469.
Dixon, L. J., R. L. Schlub, K. Pernezny, and L. E. Datnof. 2009. Host specialization of Corynespora
cassiicola. Phytopathology 99:1015–1027.
Dwyer, J. 2011. U.K. land use futures: Policy influence and challenges for the coming decades. Land
Use Policy 28(4):674–683. doi:10.1016/j.landusepol.2010.12.002.
Eastburn, D. M., M. M. Degennaro, E. H. Delucial, O. Demody, and A. J. McElrone. 2009. Elevated
atmospheric carbon dioxide and ozone alter soybean diseases at SoyFACE. Global Change
Biology 16:320–330.
Ellis, E. C., K. K. Goldewijk, S. Siebert, D. Lightman, and N. Ramankutty. 2010. Anthropogenic
transformation of the biomes, 1700 to 2000. Global Ecology & Biogeography 19:586–606.
Ersek, T., and O. K. Ribeiro. 2010. An annotated list of new Phytophthora species described post
1996. Global Change Biology Acta Phytopathologica et Entomologica Hungarica 45:251–266.
Essl, F., S. Dullinger, W. Rabitsch, P. E. Hulme, K. Hulber, V. Jarosik, I. Kleinbauer, F. Krausmann,
I. Kuhn, W. Nentwig, M. Vila, P. Genovesi, F. Gherardi, M.-L. Desprez-Loustau, A. Roques,
and P. Pysek. 2011. Socioeconomic legacy yields an invasion debt. Proceedings of the National
Academy of Sciences, USA 108:203–207.
Evans, A. M., and A. J. Finkral. 2010. A new look at spread rates of exotic diseases in North American
forests. Forest Science 56:453–459.
Fitt, B. D. L., N. Evans, P. Gladders, D. J. Hughes, J. W. Madwick, M. J. Jeger, J. A. Townsend, J.
A. Turner, and J. S. West. 2010. Climate change and arable crop disease control; mitigation and
adaptation. Proceedings of the 16th International Reinhardsbrunn Symposium, April 25–29.
In press.
Fitt, B. D. L., B. A. Fraaije, P. Chandramohan, and M. W. Shaw. 2011. Impacts of changing air com-
position on severity of arable crop disease epidemics. Plant Pathology 60:44–53.
Fleischmann, F., S. Raidl, and W. F. Osswald. 2010. Changes in susceptibility of beech (Fagus syl-
vatica) seedlings towards Phytophthora citricola under the influence of elevated atmospheric
CO2 and nitrogen fertilization. Environmental Pollution 158:1051–1060.
Frank, K. L., B. W. Geils, L. S. Kalkstein, and H. W. Thistle. 2008. Synoptic climatology of the
long-distance dispersal of white pine blister rust II. Combination of surface and upper-level
conditions. International Journal of Biometeorology 57:653–666.
APPENDIX A 291
Frenkel, O., T. L. Peever, M. I. Chilvers, H. Ozkilinc, C. Can, S. Abbo, D. Shtienberg, and A. Sherman.
2010. Ecological genetic divergence of the fungal pathogen Didymella rabiei on sympatric wild
and domesticated Cicer spp. (Chickpea). Applied and Environmental Microbiology 76:30–39.
Fry, W. E., and S. B. Goodwin. 1997. Resurgence of the Irish potato famine fungus. Bioscience
47:363–370.
Gange, A. C., E. G. Gange, A. B. Mohammad, and L. Boddy. 2011. Host shifts in fungi caused by
climate change? Fungal Ecology 4:184–190.
Garbelotto, M., R. Linzer, G. Nicolotti, and P. Gonthier. 2010. Comparing the influences of ecologi-
cal and evolutionary factors on the successful invasion of a fungal forest pathogen. Biological
Invasions 12:943–957.
Ghini, R., E. Hamada, and W. Bettiol. 2008. Climate change and plant diseases. Scientia Agricola
65:98–107.
Ghini, R., W. Bettiol, and E. Hamada. 2011. Diseases in tropical and plantation crops as affected by
climate changes: Current knowledge and perspectives. Plant Pathology 60:122–132.
Gilbert, G. S., and C. O. Webb. 2007. Phylogenetic signal in plant pathogen—host range. Proceedings
of the National Academy of Sciences, USA 104:4979–4983.
Giraud, T., P. Gladieux, and S. Gavrilets. 2010. Linking the emergence of fungal plant diseases with
ecological speciation. Trends in Ecology & Evolution 25:387–395.
Gladieux, P., E. J. Byrnes, G. Aguiletta, M. C. Fisher, J. Heitman, and T. Giraud. 2011. Epidemiology
and evolution of fungal pathogens in plants and animals. In Genetics and evolution of infectious
diseases, edited by M. Tibayrenc. Elsevier Inc. Pp. 59–132.
Goellner, K., M. Loehrer, C. Langenbach, U. Conrath, E. Koch, and U. Schaffrath. 2010. Phakopsora
pachyrhizi, the causal agent of Asian soybean rust. Molecular Plant Pathology 11:169–177.
Gonthier, P., G. Nicolotti, R. Linzer, F. Guglielmo, and M. Garbelotto. 2007. Invasion of European
pine stands by a North American forest pathogen and its hybridization with a native interfertile
taxon. Molecular Ecology 16:1389–1400.
Grulke, N. E. 2011. The nexus of host and pathogen phenology: Understanding the disease triangle
with climate change. New Phytologist 189:8–11.
Guillemaud, T., M. Ciosi, E. Lombaert, and A. Estoup. 2011. Biological invasions in agricultural set-
tings: Insights from evolutionary biology and population genetics. Comptes Rendus Biologies
334:237–246.
Haq, M., M. A. T. Mia, M. F. Rabbi, and M. A. Ali. 2011. Incidence and severity of rice diseases and
insect pests in relation to climate change. In Climate change and food security in South Asia,
edited by R. Lal, M. V. K. Sivakumar, S. M. A. Faiz, A. H. M. M. Rahman, and K. R. Islam.
Berlin, Germany: Springer. Pp. 445–457.
Harwood, T. D., X. Xu, M. Pautasso, M. J. Jeger, and M. W. Shaw. 2009. Epidemiological risk assess-
ment using linked network and grid based modelling: Phytophthora ramorum and P. kernoviae
in the U.K. Ecological Modelling 220(23):3353–3361.
Hindorf, H. 1998. Current diseases of Coffea arabica and C. canephora in East Africa causing crop
losses. Proceedings of the International Symposium on Crop Protection 50:861–865.
Holdenrieder, O., M. Pautasso, P. J. Weisberg, and D. Lonsdale. 2004. Tree diseases and landscape
processes: The challenge of landscape pathology. Trends in Ecology & Evolution 19:446–452.
Holzmüller, E. J., S. Jose, and M. A. Jenkins. 2010. Ecological consequences of an exotic fungal
disease in eastern U.S. hardwood forests. Forest Ecology and Management 259:1347–1353.
Horvath, Z. 2003. Damage in corn production and in hybrid multiplication caused by species of the
Anthicidae (Coleoptera) family. Cereal Research Communications 31:421–427.
Hovmøller, M. S., A. H. Yahyaoui, E. A. Milus, and A. F. Justesen. 2008. Rapid global spread of two
aggressive strains of a wheat rust fungus. Molecular Ecology 17:3818–3826.
Hunt, R. S., B. W. Geils, and K. E. Hummer. 2010. White pines, Ribes and blister rust: Integration
and action. Forest Pathology 40:402–417.
292 FUNGAL DISEASES
MacLeod, A., M. Pautasso, M. J. Jeger, and R. Haines-Young. 2010. Evolution of the international
regulation of plant pests and challenges for future plant health. Food Security 2:49–70.
Madden, L. V., and M. Wheelis. 2003. The threat of plant pathogens as weapons against U.S. crops.
Annual Review of Phytopathology 41:155–176.
Magyar, D., G. Frenguelli, E. Bricchi, E. Tedeschini, P. Csontos, D.-W. Li, and J. Bobvos. 2009. The
biodiversity of air spora in an Italian vineyard. Aerobiologia 25:99–109.
Marçais, B., M. Kavkova, and M.-L. Desprez-Loustau. 2009. Phenotypic variation in the phenology
of ascospore production between European populations of oak powdery mildew. Annals of
Forest Science 66:814.
Matesanz, S., A. Escudero, and F. Valladares. 2009. Impact of three global change drivers on a Medi-
terranean shrub. Ecology 90:2609–2621.
McElrone, A. J., J. G. Hamilton, A. J. Krafnick, M. Aldea, R. G. Knepp, and E. H. DeLucia. 2010.
Combined effects of elevated CO2 and natural climatic variation on leaf spot diseases of redbud
and sweetgum trees. Environmental Pollution 158:108–114.
Miller, S. A., F. D. Beed, and C. Lapierre Harmon. 2009. Plant disease diagnostic capabilities and
networks. Annual Review of Phytopathology 47:15–38.
Mills, P., K. Dehnen-Schmutz, B. Ilbery, M. Jeger, G. Jones, R. Little, A. MacLeod, S. Parker, M.
Pautasso, S. Pietravalle, and D. Maye. 2011. Integrating natural and social science perspec-
tives on plant disease risk, management and policy formulation. Philosophical Transactions B
366:2035–2044.
Mitchell, C. E., and A. G. Power. 2003. Release of invasive plants from fungal and viral pathogens.
Nature 421:625–627.
Money, N. P. 2007. The triumph of fungi: A rotten history. Oxford: Oxford University Press.
Montarry, J., D. Andrivon, I. Glais, R. Corbiere, G. Mialdea, and F. Delmotte. 2010. Microsatellite
markers reveal two admixed genetic groups and an ongoing displacement within the French
population of the invasive plant pathogen Phytophthora infestans. Molecular Ecology 19:
1965–1977.
Mordecai, E. 2011. Pathogen impacts on plant communities: Unifying theory, concepts, and empirical
work. Ecological Monographs. In press. doi:10.1890/10–2241.1
Moslonka-Lefebvre, M., M. Pautasso, and M. J. Jeger. 2009. Disease spread in small-size directed
networks: Epidemic threshold, correlation between links to and from nodes, and clustering.
Journal of Theoretical Biology 260:402–411.
Moslonka-Lefebvre, M., A. Finley, I. Dorigatti, K. Dehnen-Schmutz, T. Harwood, M. J. Jeger, X. M.
Xu, O. Holdenrieder, and M. Pautasso. 2011. Networks in plant epidemiology: From genes to
landscapes, countries and continents. Phytopathology 101:392–403.
Mundt, C. C., K. E. Sackett, L. D. Wallace, C. Cowger, and J. P. Dudley. 2009. Long-distance dis-
persal and accelerating waves of disease: Empirical relationships. The American Naturalist
73:456–466.
Newcombe, G., and F. M. Dugan. 2010. Fungal pathogens of plants in the homogocene. In Molecular
identification of fungi, edited by Y. Gherbawy and K. Voigt. Berlin, Germany: Springer. Pp.
3–34.
Newton, A. C., T. Akar, J. P. Baresel, P. J. Bebbeli, E. Bettencourt, K. V. Bladenopoulos, J. H.
Czembor, D. A. Fasoula, A. Katsiotis, K. Koutis, M. Koutsika-Sotiriou, G. Kovacs, H. Larsson,
M. A. A. P. de Carvalho, D. Rubiales, J. Russell, T. M. M. Dos Santos, and M. C. V. Patto.
2010. Cereal landraces for sustainable agriculture. Agronomy for Sustainable Development
30:237–269.
Norton, G., and M. Taylor. 2010. What pest is that? Recent developments in digital pest diagnostics.
Outlooks on Pest Management 21:236–238.
Olesen, J. E., M. Trnka, K. C. Kersebaum, A. O. Skjelvag, B. Seguin, P. Peltonen-Sainio, F. Rossi, J.
Kozyra, and F. Micale. 2011. Impacts and adaptation of European crop production systems to
climate change. European Journal of Agronomy 34:96–112.
294 FUNGAL DISEASES
Orwig, D. A. 2002. Ecosystems to regional impacts of introduced pests and pathogens: Historical
context, questions and issues. Journal of Biogeography 29:1471–1474.
Ostry, M. E., and G. Laflamme. 2009. Fungi and diseases—natural components of healthy forests.
Botany 87:22–25.
Paajanen, R., R. Julkunen-Tiitto, L. Nybakken, M. Petrelius, R. Tegelberg, J. Pusenius, M. Rousi, and
S. Kellomäki. 2011. Dark-leaved willow (Salix myrsinifolia) is resistant to three-factor (elevated
CO2, temperature and UV-B-radiation) climate change. New Phytologist 190:161–168.
Padmanabhan, S. Y. 1973. The great Bengal famine. Annual Review of Phytopathology 11:11–24.
Palm, M. E. 1999. Mycology and world trade: A view from the front line. Mycologia 91:1–12.
Parks, C. G., and P. Bernier. 2010. Adaptation of forests and forest management to changing climate
with emphasis on forest health: A review of science, policies and practices. Forest Ecology and
Management 259:657–659.
Pautasso, M. 2009. Geographical genetics and the conservation of forest trees. Perspectives in Plant
Ecology, Evolution and Systematics 11:157–189.
Pautasso, M., O. Holdenrieder, and J. Stenlid. 2005. Susceptibility to fungal pathogens of forests dif-
fering in tree diversity. In Forest diversity and function: Temperate and boreal systems, edited by
M. Scherer-Lorenzen, C. Koerner, and E. D. Schulze. Berlin, Germany: Springer. Pp. 263–289.
Pautasso, M., K. Dehnen-Schmutz, O. Holdenrieder, S. Pietravalle, N. Salama, M. J. Jeger, E. Lange,
and S. Hehl-Lange. 2010a. Plant health and global change—some implications for landscape
management. Biological Reviews 85:729–755.
Pautasso, M., M. Moslonka-Lefebvre, and M. J. Jeger. 2010b. The number of links to and from
the starting node as a predictor of epidemic size in small-size directed networks. Ecological
Complexity 7:424–432.
Pautasso, M., X.-M. Xu, M. J. Jeger, T. D. Harwood, M. Moslonka-Lefebvre, and L. Pellis. 2010c.
Disease spread in small-size directed trade networks: The role of hierarchical categories. Journal
of Applied Ecology 47:1300–1309.
Peltzer, D. A., R. B. Allen, G. M. Lovett, D. Whitehead, and D. A. Wardle. 2010. Effects of biological
invasions on forest carbon sequestration. Global Change Biology 16:732–746.
Perrings, C., S. Burgiel, M. Lonsdale, H. Mooney, and M. Williamson. 2010. International coopera-
tion in the solution to trade-related invasive species risks. Annals of the New York Academy of
Sciences 1195:198–212.
Planinšek, Š., L. Finér, A. del Campo, J. Alcazar, C. Vega-García, D. Dimitrov, and J. Capuliak.
2011. Adjustment of forest management strategies to changing climate. Ecological Studies
212:313–329.
Pugliese, M., M. L. Gullino, and A. Garibaldi. 2010. Effects of elevated CO 2 and temperature on
interactions of grapevine and powdery mildew: First results under phytotron conditions. Journal
for Plant Diseases and Plant Protection 117:9–14.
Rivas, G. G., M. F. Zapater, C. Abadie, and J. Carlier. 2004. Founder effects and stochastic dispersal
at the continental scale of the fungal pathogen of bananas Mycosphaerella fijiensis. Molecular
Ecology 13:471–482.
Roos, J., R. Hopkins, A. Kvarnheden, and C. Dixelius. 2011. The impact of global warming on plant
diseases and insect vectors in Sweden. European Journal of Plant Pathology 129:9–19.
Roper, M., A. Seminara, M. M. Band, A. Cobb, H. R. Dillard, and A. Pringle. 2010. Dispersal of
fungal spores on a cooperatively generated wind. Proceedings of the National Academy of Sci-
ences, USA 107:17474–17479.
Russell, R., M. Paterson, and N. Lima. 2009. How will climate change affect mycotoxins in food?
Food Research International 43:1902–1914.
Saleh, A. A., H. U. Ahmed, T. C. Todd, S. E. Travers, K. A. Zeller, J. F. Leslie, and K. A. Garrett.
2010. Relatedness of Macrophomina phaseolina isolates from tallgrass prairie, maize, soybean
and sorghum. Molecular Ecology 19:79–91.
APPENDIX A 295
Savage, D., M. J. Barbetti, W. J. MacLeod, M. U. Salam, and M. Renton. 2010. Timing of propagule
release significantly alters the deposition area of resulting aerial dispersal. Diversity and Dis-
tributions 16:288–299.
Savary, S., A. Mila, L. Willocquet, P. Esker, O. Carisse, and N. McRoberts. 2011. Risk factors for
crop health under global change and agricultural shifts: A framework of analyses using rice in
tropical and subtropical Asia as a model. Phytopathology 101:696–709.
Scherrer, D., and C. Körner. 2011. Topographically controlled thermal-habitat differentiation buffers
alpine plant diversity against climate warming. Journal of Biogeography 38:406–416.
Schulz, N. B. 2004. The transport system and society’s metabolism in the U.K. Population and En-
vironment 26:133–155.
Scirè, M., E. Motta, and L. D’Amico. 2011. Behaviour of Heterobasidion annosum and H. irregulare
isolates from central Italy in inoculated Pinus pinea seedlings. Mycological Progress 10:85–91.
Shaw, M. W., and T. M. Osborne. 2011. Geographic distribution of plant pathogens in response to
climate change. Plant Pathology 60:31–43.
Shin, J.-W., and S.-C. Yun. 2010. Elevated CO2 and temperature effects on the incidence of four major
chilli pepper diseases. Plant Pathology Journal 26:178–184.
Singh, D. P., D. Backhouse, and P. Kristiansen 2009. Interactions of temperature and water potential
in displacement of Fusarium pseudograminearum from cereal residues by fungal antagonists.
Biological Control 48:188–195.
Singh, S., S. Davey, and M. Cole. 2010. Implications of climate change for forests, vegetation and
carbon in Australia. New Zealand Journal of Forestry Science 40:141–152.
Skelsey, P., W. A. H. Rossing, G. J. T. Kessel, and W. van der Werf. 2009. Scenario approach for
assessing the utility of dispersal information in decision support for aerially spread plant patho-
gens, applied to Phytophthora infestans. Phytopathology 99:887–895.
Spijkerboer, H. P., J. E. Benier, D. Jaspers, H. J. Schouten, J. Goudriaan, R. Rabbinge, and W. van der
Werf. 2002. Ability of the Gaussian plume model to predict and describe spore dispersal over a
potato crop. Ecological Modelling 155:1–18.
Stack, J. P. 2010. Diagnostic networks for plant biosecurity. In Knowledge and technology transfer for
plant pathology, edited by N. Hardwick and M. L. Gullino. Dordrecht, Netherlands: Springer.
Pp. 59–74.
Stack, J. P., F. Suffert, and M. L. Gullino. 2010. Bioterrorism: A threat to plant biosecurity? In The role
of plant pathology in food safety and food security, edited by R. N. Strange and M. L. Gullino.
Dordrecht, Netherlands: Springer. Pp. 115–123.
Stewart, A., and M. Cromey. 2010. Identifying disease threats and management practices for bio-
energy crops. Current Opinion in Environmental Sustainability 3:75–80.
Stonard, J. F., A. O. Latunde-Dada, Y.-J. Huang, J. S. West, N. Evans, and B. D. L. Fitt. 2010. Geo-
graphic variation in severity of phoma stem canker and Leptosphaeria maculans/L. biglobosa
populations on U.K. winter oilseed rape (Brassica napus). European Journal of Plant Pathology
126:97–109.
Sturrock, R. N., S. J. Frankel, A. V. Brown, P. E. Hennon, J. T. Kliejunas, K. E. Lewis, J. J. Worrall,
and A. J. Woods. 2011. Climate change and forest diseases. Plant Pathology 60:133–149.
Techy, L., D. G. Schmale, and C. A. Woolsey. 2010. Coordinated aerobiological sampling of a plant
pathogen in the lower atmosphere using two autonomous unmanned aerial vehicles. Journal of
Field Robotics 27:335–343.
Thomas, K. 2010. Climate change and management of cool season grain legume crops. In Impact
of climate change on diseases of cool season grain legume crops, edited by S. S. Yadav, D. L.
McNeil, R. Redden, and S.A. Patil. Berlin, Germany: Springer. Pp. 99–113.
Thomson, L. J., S. MacFadyen, and A. A. Hoffmann. 2009. Predicting the effects of climate change
on natural enemies of agricultural pests. Biological Control 52:296–396.
Tiley, G. E. D. 2010. Biological flora of the British Isles: Cirsium arvense (L.) Scop. Journal of
Ecology 98:938–983.
296 FUNGAL DISEASES
Tubby, K. V., and J. F. Webber. 2010. Pests and diseases threatening urban trees under a changing
climate. Forestry 81:451–459.
Turnbull, L. A., J. M. Levine, A. J. F. Fergus, and J. S. Petermann. 2010. Species diversity reduces
invasion success in pathogen-regulated communities. Oikos 119:1040–1046.
Tylianakis, J. M., R. K. Didham, J. Bascompte, and D. A. Wardle. 2008. Global change and species
interactions in terrestrial ecosystems. Ecology Letters 11:1351–1363.
Woods, A. J., D. Heppner, H. H. Kope, J. Burleigh, and L. Maclauchlan. 2010. Forest health and
climate change: A British Columbia perspective. The Forestry Chronicle 86:412–422.
Worrall, J. J., G. C. Adams, and S. C. Tharp. 2010. Summer heat and an epidemic of cytospora canker
of Alnus. Canadian Journal of Plant Pathology 32:376–386.
Xu, X. M., P. Nicholson, M. A. Thomsett, D. Simpson, B. M. Cooke, F. M. Doohan, J. Brennan, S.
Monaghan, A. Moretti, G. Mule, L. Hornok, E. Beki, J. Tatnell, A. Ritieni, and S. G. Edwards.
2008. Relationship between the fungal complex causing Fusarium head blight of wheat and
environmental conditions. Phytopathology 98:69–78.
Xu, X. M., T. D. Harwood, M. Pautasso, and M. J. Jeger. 2009. Spatio-temporal analysis of an inva-
sive plant pathogen (Phytophthora ramorum) in England and Wales. Ecography 32:504–516.
Young, J. M., C. Allen, T. Coutinho, T. Denny, J. Elphinstone, M. Fegan, M. Gillings, T. R. Gottwald,
J. H. Graham, N. S. Iacobellis, J. D. Janse, M. A. Jacques, M. M. Lopez, C. E. Morris, N.
Parkinson, P. Prior, O. Pruvost, J. Rodrigues Neto, M. Scortichini, Y. Takikawa Y, and C. D.
Upper. 2008. Plant–pathogenic bacteria as biological weapons—real threats? Phytopathology
98:1060–1065
Zocca, A., C. Zanini, A. Aimi, G. Frigimelica, N. La Porta, and A. Battisti. 2008. Spread of plant
pathogens and insect vectors at the northern range margin of cypress in Italy. Acta Oecologica
33:307–313.
A15
Summary
Several fungal diseases of non-domestic animal species have been described
as agents of epizootic proportions in wild animals in recent years. The emergence
of these diseases has reshaped the understanding of the role of fungi as contagious
diseases having an impact on wild animal populations. The recent description of
fungal epizootics caused by Batrachochytrium dendrobatidis (Bd) in amphibians
worldwide and Geomyces destructans in North American bats has called attention
to the factors driving the emergence of these diseases. Two other fungal diseases
of wild animals, lobomycosis in dolphins and penicilliosis in wild bamboo rats,
have been recognized for their zoonotic potential and highlight the need for
comprehensive, multidisciplinary approaches to understanding the ecology and
epidemiology of these diseases in wild animal populations. Continued efforts
aimed at mitigating the effects of fungal epizootics on wildlife populations and on
public health will only be successful through the identification of factors driving
the emergence of these diseases across different taxa, and embracing the concept
that wildlife may serve as sentinels of ecosystem health.
Introduction
Anthropogenic activities have been the likely driving factors behind the
emergence of some diseases in wildlife (Daszak et al., 2000). In some cases,
animal species whose status may have been threatened by other factors may now
be faced with extinction as the emergent disease spreads through a diminished
population. The recent spread of two major fungal epizootic agents, namely
Batrachochytrium dendrobatidis and Geomyces destructans in amphibians and
bats, respectively, could result in the largest changes to vertebrate populations
in recorded history (Frick et al., 2010; Skerrat et al., 2007). The discovery of
these fungal agents as major causes of wildlife epizootics in the past decade has
revolutionized the way that biologists approach the detection and diagnosis of
fungal diseases, challenging the prior misconception that fungal infections only
occurred “sporadically or in small outbreaks” and were more important to captive
wildlife, where captivity was thought to “increase susceptibility to these diseases”
(Burek, 2001).
As novel pathogens have been discovered, or known ones recognized in
novel hosts, in novel geographic areas, or with increased incidence, wildlife
conservationists have been faced with the emergence of fungal diseases that
threaten the status of wild animal populations. The factors driving disease in
free-ranging wild animal species can no longer be easily differentiated from
those affecting captive wildlife. The different factors have been blurred into a
continuum of factors through modern globalization, inadvertent movement of
disease, disease vectors, or animals themselves, through the trade of animals
across geographical barriers and the mixing of potential hosts of disease that
may not have been exposed to each other in their native habitats. Many of these
anthropogenic actions have expanded the geographic range of some diseases
or removed the natural barriers that had prevented their spread, exposing naïve
hosts to pathogens to which they were not previously exposed. Concurrently,
many of the environmental factors thought to predispose their captive counter-
parts to infectious diseases have been identified and minimized through modern
captive animal science aimed at reducing stress levels, providing better environ-
mental conditions, and through better quarantine, improving disease screening
and recognition procedures.
Institutions dedicated to the captive care of animals for conservation pur-
poses often encounter infectious diseases that would otherwise go unreported
in those species. These provide a useful baseline of understanding the host–
pathogen relationship systems. Captive animal populations can serve as viable
models for recognizing or refining the understanding of disease mechanisms
in individual hosts, which may not be possible in free-ranging animals. For in-
298 FUNGAL DISEASES
Batrachochytrium dendrobatidis
Batrachochytrium dendrobatidis is a member of a basal group of fungi, the
Chytridiomycota, and the only member known to affect vertebrates. Bd infects
amphibian species, with frogs most often reported. This emerging pathogen was
discovered in the late 1990s (Berger et al., 1998; Longcore et al., 1999; Pessier
et al., 1999) and has been recognized as a significant pathogen partially impli-
cated in the global decline of amphibian populations (Schloegel et al., 2006;
Skerrat et al., 2007). It has been retrospectively identified in North American
amphibians as early as 1961 (Ouellet et al., 2005). Bd may be the most signifi-
cant fungal infectious disease agent of vertebrate species, based on the global
distribution, wide species host range, pathogenic potential, and ability to cause
large-scale mortality.
Molecular genetic evidence from isolates collected from different locations
worldwide suggests that recent Bd outbreaks were caused by a pathogen recently
disseminated (Morehouse et al., 2003), and anthropogenic spread is suspected
APPENDIX A 299
in at least some introductions (Weldon et al., 2004), but the origin is still not
conclusively known. Virulence and pathogenic potential varies among isolates,
and some may show little or no pathogenicity to their hosts (Berger et al., 2005;
Goka et al., 2009), but these host-adapted strains may still serve as reservoirs
of strains that are pathogenic to other species. The naturally occurring bacterial
flora of the host may produce antifungal proteins that decrease the ability of the
fungi to colonize amphibian skin, and these bacteria may improve host survival
(Harris et al., 2006). Mortality in affected amphibians is likely associated with
physiologic abnormalities (electrolyte loss and osmotic imbalances) caused by
damage to the permeable amphibian skin.
Clinical signs in affected frogs can be subtle and not detectable before sud-
den death (Pessier et al., 1999). Affected frogs show excessive shedding of skin,
usually on the legs, feet, and ventrum (Nichols et al., 2001). Molecular techniques
have been established as a cost-effective and rapid method to detect the organism
in amphibian samples (Boyle et al., 2004; Kriger et al., 2006). Histopathology,
showing intralesional organisms in keratinized layers of skin, and cytology of
shed skin or imprints showing chytrid organisms have been used for diagnosing
infections (Nichols et al., 2001). Bd has no known zoonotic potential.
Geomyces destructans
Geomyces destructans is a newly described psychrophilic fungus that affects
North American bats during hibernation (Gargas et al., 2009), causing the char-
acteristic fungal white growth after which the syndrome was originally named.
White-nose syndrome (WNS) could more accurately be dubbed bat geo-
mycosis (per Chaturvedi and Chaturvedi, 2011), but the denomination of WNS
has been used since its recognition (Blehert et al., 2009). Its impact on wild bat
populations has earned it the distinction of being the second most significant
vertebrate pathogen in recorded history (after Bd), if only because Bd has been
documented in more species, with a wider global impact.
Bat geomycosis is a true emerging disease of epizootic proportion, appearing
in upstate New York in 2006 (Blehert et al., 2009), but quickly spreading south
and westward and devastating bat populations (Blehert et al., 2009; Frick et al.,
2010). Host susceptibility to infection seems to vary with species. The fungus has
also been identified in Europe (Puechmaille et al., 2010; Wibbelt et al., 2010),
but to date, the severe, widespread mortality seen in North American bats has not
been documented. Among North American bat species, either pathologic lesions
or G. destructans DNA has been detected in the gray bat (Myotis grisescens),
the Indiana bat (M. sodalis), the little brown bat (Myotis lucifugus), the northern
long-eared bat (M. septentrionalis), the eastern small-footed bat (M. leibii), the
southeastern bat (M. austroriparius), the cave bat (M. velifer), the tricolored bat
APPENDIX A 301
(Perimyotis subflavus), and the big brown bat (Eptesicus fuscus) (Foley et al.,
2011).
Geomyces destructans infections in bats result in premature arousal from
hibernation and abnormal behavior. Although the mechanisms by which death
occurs are still under investigation, the inability to forage in the winter after pre-
mature arousal (due to lack of prey availability) and direct damage to the wing
membranes, resulting in irreversible physiologic and homeostatic deficits from
which bats cannot recover (Cryan et al., 2010), are significant factors.
Limiting human access to caves of concern has been a management tool
implemented to limit the spread of the Geomyces fungus. One of the first theories
to explain the rapid appearance and spread of this pathogen in North American
bat populations suggested that the fungus may have been endemically established
in European bat populations and that a recent anthropogenic introduction into
North America may have led to disease in naïve populations, but this theory
has not been tested. Histopathology of rostral muzzle and wing membranes is
deemed important to confirm infections and establish true prevalence of this dis-
ease (Meteyer et al., 2009), although advanced diagnostic tools are likely being
developed and refined.
As bats are significant providers of ecological services, such as insect control
and pollination, the extinction or even reduction of bat populations is likely to
have economic impacts on society beyond the immediate loss of biodiversity.
Penicillium marneffei
Penicillium marneffei is the only fungus of this genus known to be a primary
pathogen of mammals. Penicilliosis marneffei is a systemic fungal disease of wild
rodents and humans recognized in northeast India and Southeast Asia (Thailand,
the Guangxi region of China, Vietnam, Taiwan, and Hong Kong). Penicillium
marneffei was first identified from hepatic lesions in a bamboo rat (Rhizomys
sinensis) from Dalat, South Vietnam (Caponi et al., 1956), and was subsequently
identified as a human pathogen following an accidental exposure by a researcher
(Vanittanakom et al., 2006). It is an emerging human disease, a primary pathogen
to bamboo rats, and a threat to public health. Penicilliosis marneffei is third only
to tuberculosis and cryptococcosis as the most common opportunistic infections
in patients with AIDS in northern Thailand (Vanittanakom et al., 2006), and the
source of more than 100 cases per year in the Guanxi region of China (Cao et al.,
2011).
Affected rodents show ascites and enlargement of the liver, spleen, and
lymph nodes. The fungal pathogens can be identified in multiple organs and
ascitic fluid, but is most commonly cultured from the lung of affected rodents
(Ajello et al., 1995).
Wild bamboo rats may be good sentinels of this sapronotic disease, and are
significant in the ecology of this disease, although their exact role is not com-
302 FUNGAL DISEASES
pletely understood. Penicillium marneffei has been isolated from the internal
organs of four species of bamboo rats (Rhizomys sinensis, Rhizomys pruinosus,
Rhizomys sumatrensis, and Cannomys badius) and from the soil associated with
their burrows (Vanittanakom et al., 2006). A high prevalence of infection and le-
sions among wild bamboo rats of the genera Rhizomys and Cannomys suggested
that these wild animals may serve as enzootic reservoirs, but the prevalence var-
ies across regions, and a wildlife reservoir has never been conclusively identified
(Cao et al., 2011; Deng et al., 1986; Gugnani et al., 2004; Vanittanakom et al.,
2006). When sympatric rodents have been sampled, only bamboo rats appear to
harbor the organism (Gugnani et al., 2004), suggesting distinct host differences.
Possibly, bamboo rats are exposed from either an unidentified wildlife vector
or an environmental source (Vanittanakom et al., 2006), and are not the wildlife
hosts of this disease, although they can transmit the disease to humans. The
role of wild bamboo rats as amplifiers or dispersers of infectious stages has not
been eliminated. The initial case of human infection in 1959 in a researcher who
became infected after injection from a needle used in laboratory inoculations
(Vanittanakom et al., 2006) underscores the zoonotic risk posed by wildlife.
Although most human cases involve immunosuppressed individuals, infec-
tions in humans with normal immunity—and the lack of evidence for immu-
nosuppression in affected bamboo rats—suggests that P. marneffei could be a
primary mammalian fungal pathogen (Duong, 1996). A complete understanding
of the role of wildlife in this disease and its dynamics in rodent populations is
essential to better managing the threat to human health in endemic regions and
mitigating the possible anthropogenic spread of this disease through travel, move-
ment of commercial products (some of which may be environmental reservoirs),
and the primary trade of wild animals. Penicilliosis marneffei remains one of the
most enigmatic emerging fungal diseases of wildlife, and one whose primary
zoonotic potential is not fully understood.
Lacazia loboi
Lobomycosis is a zoonotic disease of dolphins caused by Lacazia loboi, a
cutaneous fungus in the order Onygenales that has never been cultured (Herr
et al., 2001). The disease affects humans and dolphins in tropical and transitional
tropical climates. Lobomycosis is endemic in certain human populations in Cen-
tral and South America, and is likely endemic in regional dolphin populations
(Murdoch et al., 2008), although the overall prevalence and many factors of its
ecology are still unclear. The disease has been confirmed in two dolphin species,
Guiana dolphins (Sotalia guianensis) (Van Bressem et al., 2009) and Atlantic
bottlenose dolphins (Tursiops truncatus) (Caldwell et al., 1975), and has been
suspected in Indo-Pacific bottlenose dolphins (T. aduncus) (Kiszka et al., 2009).
Transmission between dolphins is likely by direct contact, and affected
dolphins show chronic white to pink verrucous, raised lesions that may coalesce
APPENDIX A 303
into large plaques or nodules predominantly on dorsal and pectoral fins, the head,
fluke, and caudal peduncle (Murdoch et al., 2010). Lesions may be associated
with sites of prior trauma, and many affected dolphins have shown impaired
adaptive immunity, suggesting that the disease may represent an opportunistic
infection in an immunocompromised host (Reif et al., 2008).
Some have suggested that the disease in dolphins is being more frequently
reported (Van Bressem et al., 2009) and the reported geographic range may be
expanding (Rotstein et al., 2009), making it a true emerging fungal disease of
wildlife. In the Indian River lagoon of Florida, measured prevalence was 30 per-
cent in Atlantic bottlenose dolphins (Tursiops truncatus) from the southern part of
the lagoon in a 2006 survey (Reif et al., 2006). It was not detected in the northern
portion (near Charleston, South Carolina), suggesting that localized geographical
factors or environmental stressors may play a significant role in the incidence and
distribution of the disease. A subsequent study (Murdoch et al., 2008) suggested
that the disease is endemic and not an emerging disease in the Indian River dol-
phin population. A survey of Guiana dolphins (Van Bressem et al., 2009) in the
Paranaguá estuary in Brazil suggested an increasing detection of a lobomycosis-
like disease (missing histological confirmation) and suggested that the change
was an indication of the health of the marine environment.
Conclusion
Several emerging fungal diseases have recently been shown to affect wild
animal species. Some of these diseases carry a zoonotic risk potential, but even
those that do not directly affect human health are likely to carry a societal cost in
terms of ecosystem health. Wild animal species can be sentinels of emerging dis-
eases and are early indicators of overall ecosystem health. Advances in veterinary
medicine have aided in the recognition of fungal pathogens, and new techniques
are being developed to mitigate their impact on wild populations. However, the
inherent responsibility falls to veterinarians, physicians, conservation biologists,
and public health professionals to properly document and disseminate the find-
ings of fungal diseases in novel hosts, geographic locations, or areas of increased
frequency. Continued multidisciplinary surveillance of fungal diseases is es-
sential to understand the impact of these emerging pathogens on wild animal
populations, humans, and ecosystems.
APPENDIX A 309
References
AARK (Amphibian Ark). 2011. Amphibian Ark: About us. http://www.amphibianark.org/about-us/
(accessed February 27, 2011).
Abarca, M. L., J. Martorell, G. Castellá, A. Ramis, and F. J. Cabañes. 2008. Cutaneous hyalohy-
phomycosis caused by a Chrysosporium species related to Nannizziopsis vriesii in two green
iguanas (Iguana iguana). Medical Mycology 46:349–354.
Ajello, L., A. A. Padhye, S. Sukroongreung, C. H. Nilaku, and S. Tantimavanic. 1995. Occurrence of
Penicillium marneffei infections among wild bamboo rats in Thailand. Mycopathologia 131:1–8.
Berger, L., R. Speare, P. Daszak, D. E. Green, A. A. Cunningham, C. L. Goggin, R. Slocombe, M. A.
Raga, A. D. Hyatt, K. R. McDonald, H. B. Hines, K. R. Lips, G. Marantelli, and H. Parke. 1998.
Chytridiomycosis causes amphibian mortality associated with population declines in the rain
forests of Australia and Central America. Proceedings of the National Academy of Sciences,
USA 95:9031–9036.
Berger, L., G. Marantelli, L. L. Skerratt, and R. Speare. 2005. Virulence of the amphibian chytrid
fungus Batrachochytrium dendrobatidis varies with the strain. Diseases of Aquatic Organisms
68:47–50.
Bertelsen, M. F., G. J. Crawshaw, L. Sigler, and D. A. Smith. 2005. Fatal cutaneous mycosis in ten-
tacled snakes (Erpeton tentaculatum) caused by the Chrysosporium anamorph of Nannizziopsis
vriesii. Journal of Zoo and Wildlife Medicine 36:82–87.
Blehert, D. S., A. C. Hicks, M. Behr, C. U. Meteyer, B. M. Berlowski-Zier, E. L. Buckles, J. T. H.
Coleman, S. R. Darling, A. Gargas, R. Niver, J. C. Okoniewski, R. J. Rudd, and W. B. Stone.
2009. Bat white nose syndrome: An emerging fungal pathogen? Science 232:227.
Bowman, M. R., J. A. Paré, L. Sigler, J. P. Naeser, K. K. Sladky, C. S. Hanley, P. Helmer, L. A.
Phillips, A. Brower, and R. Porter. 2007. Deep fungal dermatitis in three inland bearded dragons
(Pogona vitticeps) caused by the Chrysosporium anamorph of Nannizziopsis vriesii. Medical
Mycology 45:371–376.
Boyle, D. G., D. B. Boyle, V. Olsen, J. A. T. Morgan, and A. D. Hyatt. 2004. Rapid quantitative de-
tection of chytridiomycosis (Batrachochytrium dendrobatidis) in amphibian samples using real
time Taqman PCR assay. Diseases of Aquatic Organisms 60:141–148.
Burek, K. 2001. Mycotic diseases. In Infectious diseases of wild mammals, 3rd ed., edited by E. S.
Williams and I. K. Barker. Ames, IA: Iowa State University Press. Pp. 514–531.
Caldwell, D. K., M. C. Caldwell, J. C. Woodard, L. Ajello, W. Kaplan, and H. M. McClure. 1975.
Lobomycosis as a disease of the Atlantic bottle-nosed dolphin (Tursiops truncatus, Montagu,
1821). American Journal of Tropical Medicine and Hygiene 24:105–114.
Cao, C., L. Liang, W. Wang, H. Luo, S. Huang, D. Liu, J. Xu, D. A. Henk, and M. C. Fisher. 2011.
Common reservoirs for Penicillium marneffei infection in humans and rodents, China. Emerging
Infectious Diseases 17:209–214.
Caponi, M., G. Segretain, and P. Sureau. 1956. Penicilliosis de Rhizomis sinensis. Bulletin de la
Société de Pathologie Exotique 49:418–421.
Cassone, A., and A. Torosantucci. 2006. Opportunistic fungi and fungal infections: The challenge of
a single, general antifungal vaccine. Expert Review of Vaccines 5:859–867.
Chaturvedi, V., and S. Chaturvedi. 2011. What is in a name? A proposal to use geomycosis instead
of white nose syndrome (WNS) to describe bat infection caused by Geomyces destructans.
Mycopathologia 171:231–233.
Cryan, P. M., C. Uphoff, C. U. Meteyer, J. G. Boyles, and D. S. Blehert. 2010. Wing pathology of
white-nose syndrome in bats suggests life-threatening disruption of physiology. BMC Biology
8:135–142.
Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2000. Emerging infectious diseases of wildlife—
threats to biodiversity and human health. Science 287:443–449.
Deng, Z. L., M. Yun, and L. Ajello. 1986. Human Penicilliosis marneffei and its relation to the bamboo
rat (Rhizomys pruinosus). Journal of Medical and Veterinary Mycology 24:383–389.
310 FUNGAL DISEASES
A16
Introduction
The emergence of fungal and fungal-like plant pathogens in agricultural and
natural ecosystems can be triggered by a number of key changes in the host–
pathogen–environment interaction (Anderson et al., 2004; Desprez-Loustau and
Rizzo, 2011; Desprez-Loustau et al., 2007). The evolution of pathogens owing
to selection pressure due to fungicides, deployment of resistant plant varieties, or
changes in the environment (global to local) can allow previously known patho-
gens to increase in incidence locally or across wide geographic areas (Anderson
et al., 2004). Movement of fungi from one part of the world to another may
also lead to disease emergence and large-scale epidemics of plant pathogens
(Desprez-Loustau et al., 2007; Rizzo, 2005). Because of the well-known impacts
of exotic plant pathogens, much effort has been made at regional, national, and
The Pathogen
Phytophthora ramorum was unknown before it was observed to cause dis-
eases of a number of host species in the mid-1990s in Europe and California.
The pathogen was officially named P. ramorum in 2001 (Werres et al., 2001).
Population studies indicate that P. ramorum in both North America and Europe
has a genetic structure consistent with that expected of an introduced species and
that it reproduces exclusively clonally (Ivors et al., 2004, 2006). P. ramorum has
two mating types (A1, A2); to date, all North American isolates of P. ramorum
have been found to be mating type A2, while all isolates in Europe (with rare
exceptions) are A1 (Grünwald et al., 2008). Sexual reproduction outside of the
laboratory has not been documented (Grünwald et al., 2008), although laboratory
attempts at crossing the P. ramorum mating types have produced viable progeny
with similar virulence to the parent types (Boutet et al., 2010). Three genetically
distinct lineages (NA1, NA2, and EU1) have been identified within P. ramorum
(Grünwald et al., 2008; Ivors et al., 2006). Only NA1 is present in California and
Oregon forests, all three lineages are found in North American nursery popula-
tions, and only EU1 has been found in European nurseries and woodlands (Goss
et al., 2009, 2011; Grünwald et al., 2008; Ivors et al., 2006).
The putative exotic nature of P. ramorum in North America is also supported
by the findings that it is present in nurseries (Yakabe et al., 2009) yet is absent in
historical herbarium collections (Monahan et al., 2008). However, the geographic
origin of P. ramorum remains unknown. From a phylogenetic perspective, its
314 FUNGAL DISEASES
agent of the disease in 2000 (Rizzo et al., 2002). In the United States, P. ramo-
rum’s current geographic range in native forests extends from the Big Sur area in
central California to southern Mendocino County, with two disjunct populations
in Humboldt County, California, and one small population in Curry County, Or-
egon (Figure A16-1). P. ramorum has not become established in forests outside
of this area. Potentially millions of tanoak and oak trees have been lost to the
disease over the past 10 years (Meentemeyer et al., 2008c, 2011).
The pathogen has been associated with the horticultural industry in both
the United States and Canada and has been consistently found in a number of
nurseries. While dozens of plant species have been found infected in nurseries,
the majority of infections have been associated with the genera Rhododendron,
Camellia, Pieris, Kalmia, and Viburnum (Osterbauer et al., 2004; Parke et al.,
2010a). Although P. ramorum has not been found in forests outside of California
and Oregon, the pathogen has become established in streams in several states,
including Washington, Alabama, Mississippi, Georgia, and South Carolina (Oak
et al., 2010). These streams have been associated with infestations in nearby nurs-
eries; in several rare instances vegetation along the banks has become infected by
P. ramorum, but there is no evidence of terrestrial colonization of forests at this
time outside of California and Oregon (Oak et al., 2010).
The emergence of P. ramorum in California and Oregon forests can be linked
to several key factors: movement of ornamental plants (both into and within
California) (Ivors et al., 2006; Prospero et al., 2009), the susceptibility and dis-
tribution of host plants in coastal forests (Anacker et al., 2008; Dodd et al., 2004,
2008; Hayden et al., 2011; Hüberli and Garbelotto, 2011; Meentemeyer et al.,
2008a,b,c; Rizzo et al., 2005), and a suitable climate (Meentemeyer et al., 2004,
2011). P. ramorum appears to have originally spread from two focal points in
California. Two P. ramorum populations (Marin and Santa Cruz) were identified
using population genetic analysis as the two oldest sources of the pathogen in
the state (Mascheretti et al., 2008, 2009). This genetic analysis corroborated an-
ecdotal evidence from field observations and nursery records. By reconstructing
the epidemic from this point, other locations were determined to be of intermedi-
ate age. For example, the Big Sur population was shown to be derived from the
Santa Cruz population originally associated with nursery stock (Mascheretti et al.,
2008). This analysis suggests at least eight separate introductions of P. ramorum
within California from the two focal populations. Most of these introductions
were likely human related via movement of nursery plants, but some introduc-
tions were likely caused by natural spread of the pathogen from forests (Cushman
and Meentemeyer, 2008; Mascheretti et al., 2008, 2009).
P. ramorum can be found in a number of forest types along California and
Oregon coasts; these range from drier oak mixed evergreen forests dominated by
coast live oak to wetter forest types dominated by coast redwood or Douglas fir
(Rizzo and Garbelotto, 2003). In these conifer-dominated forests, tanoak is the
316 FUNGAL DISEASES
host most affected by sudden oak death. As forest types vary across a topographi-
cal landscape, there are associated changes in microclimate and, consequently, a
likelihood of variation in the host–pathogen interaction and pathogen transmis-
sion (Condeso and Meentemeyer, 2007; Meentemeyer et al., 2008a, 2011). These
changes in microclimate may be due to variation in edaphic factors such as as-
pect, soil strata, and hydrology that underlie the growth of a particular vegetation
type. Differences in the ensuing physical structure of the vegetation itself also
affect light availability and, consequently, temperature and moisture. These are
crucial factors that determine survival and sporulation of most pathogens, includ-
ing Phytophthora species. On the level of fine-scale host-pathogen interactions,
these environmental variables may affect the timing and production of inoculum,
the length of the infectious period, the degree of host susceptibility or pathogen
virulence, pathogen survival through adverse conditions (dormancy), the timing
and incidence of new infections, and the overall magnitude and pattern of disease
epidemics.
Although P. ramorum infects more than 25 host species in these woodlands,
nonlethal foliar lesions on bay laurel (Umbellularia californica) are the most
important host tissue for sporulation by P. ramorum in the oak woodlands of
California (Davidson et al., 2005, 2008, 2011). Levels of inoculum in through-fall
rain are up to 20 times higher under bay laurel as opposed to other forest trees.
In addition, at the landscape level, infection on bay laurel is known to precede
infection on oak and tanoak trees, and the presence of this host is associated with
higher levels of oak mortality (Cobb et al., 2010; Maloney et al., 2005). Conse-
quently, infections on bay laurel leaves drive the spread of P. ramorum and the
onset of lethal infections on oak and tanoak. Sporulation does occur on tanoak
leaves and this species does appear capable of driving epidemics in the absence
of bay laurel (Davidson et al., 2008). This situation is occurring in Oregon forest,
where bay laurel is a relatively minor component of the forests (Hansen et al.,
2008).
Various modeling efforts have been made to predict the spread and estab-
lishment of P. ramorum in California, Oregon, and other locations within North
America (e.g., Kelly et al., 2007; Meentemeyer et al., 2004, 2008a,b, 2011;
Vaclavik et al., 2010; Venette and Cohen, 2006). For example, a recent epidemio-
logical model was used in combination with geographical modeling to predict
the spread of P. ramorum through host populations in wildland forests, subject
to fluctuating weather conditions (Meentemeyer et al., 2011). Application of the
model to Californian landscapes over a 40-year period (1990–2030), since the ap-
proximate time of pathogen introduction, suggests that in the absence of extensive
control, a 10-fold increase in disease spread will occur between 2010 and 2030
with most infection concentrated along the northern coast of California between
San Francisco and Oregon (Meentemeyer et al., 2011). Long-range dispersal
of inoculum to susceptible host communities in the Sierra Nevada foothills and
coastal southern California leads to little secondary infection due to lower host
318 FUNGAL DISEASES
availability and less suitable weather conditions. However, a shift to wetter and
milder conditions in future years would double the amount of disease spread
in California through 2030 (Meentemeyer et al., 2011). In other areas of North
America, the forests of the southern Appalachian Mountains are considered to
be at the highest risk (Kelly et al., 2007; Rizzo et al., 2005; Venette and Cohen,
2006). The combination of susceptible oaks (e.g., Q. rubra), potential sporulating
hosts found in the understory (Rhododendron, Kalmia), and a moist, relatively
mild climate has the potential to support the emergence of P. ramorum if it is
introduced into these areas (Spaulding and Rieske, 2011).
Disease Management
Management of P. ramorum–associated diseases in forests, woodlands, and
urban areas has taken a multiscale approach ranging from individual trees to
landscapes to international quarantines (Alexander and Lee, 2010; Frankel, 2008;
Rizzo et al., 2005). Disease prevention and mitigation at the individual plant level
or urban–wildland interface in California has been focused on chemical control or
other programs designed to maintain health of plants. Some fungicides have been
developed that act as protectants (e.g., phosphonates) against infection, but few
chemicals have been developed that work once the plant is infected (Garbelotto
and Schmidt, 2009; Garbelotto et al., 2009). Removal of inoculum-producing
plants, such as bay laurel or rhododendron, has also been important at smaller
scales to protect high-value oaks (Swiecki and Bernhardt, 2008). Education and
involvement of local communities has been critical at the urban–wildland inter-
face to the implementation of management programs (Alexander and Lee, 2010).
At larger landscape scales in wildland forest communities, management
strategies for P. ramorum have included prevention, eradication, treatment, and
restoration (Rizzo et al., 2005; Valachovic et al., 2008, 2010). Eradication has
been attempted in some cases, most notably with tanoak forests in Oregon
(Kanaskie et al., 2010) and larch plantations in the United Kingdom (Brasier
and Webber, 2010), but has met with mixed success. Important successes have
been balanced by continuing tree mortality in many areas (Kanaskie et al., 2010).
Difficulties have been encountered in detecting the pathogen at an early enough
stage for eradication to be completely effective at a landscape scale. Cryptic
infection (i.e., with minimal or no symptoms) of foliage during the initial inva-
sion of a site by P. ramorum has allowed the pathogen to stay one step ahead
of detection efforts in many cases. The development of management strategies,
beyond eradication, for forest lands following invasion by P. ramorum is still
in the early stages (Rizzo et al., 2005; Valachovic et al., 2008, 2010). Decision
making requires the ability to fit disease management into the context of other
management goals (e.g., fire prevention, wildlife) within the broader forest land-
scape (Rizzo et al., 2005). Examples of approaches that are being tested include
forest stand thinning to remove inoculum-producing hosts and use of prescribed
fire (Valachovic et al., 2008, 2010).
The broadest scale for disease management, regional to international, is
driven by regulations and management practices designed to prevent further
spread of Phytophthora (Brasier, 2008; Frankel, 2008; Rizzo et al., 2005). In
recent years, broadening of national and international quarantines designed to
prevent pathogen movement has led to an increased effort to manage all Phytoph-
thora diseases in nursery settings (Osterbauer et al., 2004; Parke et al., 2010a,b).
While dozens of plant species have been found infected in nurseries, the major-
ity of infections have been associated with the genera Rhododendron, Camellia,
Pieris, Kalmia, and Viburnum (Osterbauer et al., 2004; Parke et al., 2010a). These
plant species have become the focal point for development of best management
practices and pathogen detection strategies (Parke et al., 2010a). The need for
320 FUNGAL DISEASES
References
Alexander, J., and C. A. Lee. 2010. Lessons learned from a decade of sudden oak death in California:
Evaluating local management. Environmental Management 46:315–328.
Anacker, B. L., N. E. Rank, D. Hüberli, M. Garbelotto, S. Gordon, T. Y. Harnik, R. Whitkus, and
R. K. Meentemeyer. 2008. Susceptibility to Phytophthora ramorum in a key infectious host:
Landscape variation in host genotype, host phenotype, and environmental factors. New Phytolo-
gist 177:756–766.
Anderson, P. K., A. A. Cunningham, N. G. Patel, F. J. Morales, P. R. Epstein, and P. Daszak. 2004.
Emerging infectious diseases of plants: Pathogen pollution, climate change and agrotechnology
drivers. Trends in Ecology & Evolution 19:535–544.
Aukema, J. A., S. McCullough, B. Von Holle, A. M. Liebhold, K. Britton, and S. J. Frankel. 2010.
Historical accumulation of nonindigenous forest pests in the continental United States. BioSci-
ence 60:886–897.
Brasier, C. M. 2008. The biosecurity threat to the UK and global environment from international trade
in plants. Plant Pathology 57:792–808.
———. 2009. Phytophthora biodiversity: How many Phytophthora species are there? In Proceedings
of the fourth meeting of the International Union of Forest Research Organizations (IUFRO)
Working Party S07.02.09: Phytophthoras in forests and natural ecosystems, edited by E. M.
Goheen and S. J. Frankel. Albany, CA: U.S. Department of Agriculture, Forest Service, Pacific
Southwest Research Station. Gen. Tech. Rep. PSW-GTR-221. Pp. 101–116.
Brasier, C. M., and J. Webber. 2010. Sudden larch death. Nature 466:824–825.
Brasier, C. M., A. M. Vettraino, T. T. Chang, and A. Vannini. 2010. Phytophthora lateralis discovered
in an old growth Chamaecyparis forest in Taiwan. Plant Pathology 59:595–603.
Boutet, X., A. Vercauteren, K. Heungens, F. Laurent, and A. Chandelier. 2010. Oospores progenies
from Phytophthora ramorum. Fungal Biology 114:369–378.
Cobb, R. C., R. K. Meentemeyer, and D. M. Rizzo. 2010. Apparent competition in canopy trees
determined by transmission rather than susceptibility. Ecology 91:327–333.
Condeso, T. E., and R. K. Meentemeyer. 2007. Effects of landscape heterogeneity on the emerging
forest disease sudden oak death. Journal of Ecology 95:364–375.
Cushman, J. H., and R. K. Meentemeyer. 2008. Multi-scale patterns of human activity and the inci-
dence of an exotic forest pathogen. Journal of Ecology 96:766–776.
Davidson, J. M., S. Werres, M. Garbelotto, E. M. Hansen, and D. M. Rizzo. 2003. Sudden oak death
and associated diseases caused by Phytophthora ramorum. Online. Plant Health Progress,
doi:10.1094/PHP-2003-0707-01-DG.
APPENDIX A 321
Hüberli, D. and M. Garbelotto. 2011 (in press). Phytophthora ramorum is a generalist plant patho-
gen with differences in virulence between isolates from infectious and dead-end hosts. Forest
Pathology.
Ivors, K., K. J. Hayden, P. J. M. Bonants, D. M. Rizzo, and M. Garbelotto. 2004. AFLP and phy-
logenetic analyses of North American and European populations of Phytophthora ramorum.
Mycological Research 108:378–392.
Ivors, K., M. Garbelotto, I. D. E. Vries, C. Ruyter-Spira, B. T. E. Hekkert, N. Rosenzweig, and P. J. M.
Bonants. 2006. Microsatellite markers identify three lineages of Phytophthora ramorum in US
nurseries, yet single lineages in US forest and European nursery populations. Molecular Ecol-
ogy 15:1493–1505.
Kanaskie, A., E. M. Hansen, E. M. Goheen, N. Osterbauer, M. McWilliams, J. Laine, M. Thompson,
S. Savona, H. Timeus, B. Woosley, W Sutton, P. Reeser, R. Schultz, and D. Hilburn. 2010.
Detection and eradication of Phytophthora ramorum from Oregon forests, 2001–2008. In
Proceedings of the sudden oak death fourth science symposium, edited by S. J. Frankel, J. T.
Kliejunas, and K. M. Palmieri. Albany, CA: U.S. Department of Agriculture, Forest Service,
Pacific Southwest Research Station. Gen. Tech. Rep. PSW-GTR-229. Pp. 3–5.
Kelly, M., Q. Guo, D. Liu, and D. Shaari. 2007. Modeling the risk of a new invasive forest disease in
the United States: An evaluation of five environmental niche models. Computers, Environment
and Urban Systems 31:689–710.
Kliejunas, J. T. 2010. Sudden oak death and Phytophthora ramorum: A summary of the literature.
Albany, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Sta-
tion. 181 pp.
Maloney, P. E., S. C. Lynch, S. F. Kane, C. E. Jensen, and D. M. Rizzo. 2005. Establishment of an
emerging generalist pathogen in redwood forest communities. Journal of Ecology 93:899–905.
Mascheretti, S., P. J. P. Croucher, A.Vettraino, S. Prospero, and M. Garbelotto. 2008. Reconstruction
of the sudden oak death epidemic in California through microsatellite analysis of the pathogen
Phytophthora ramorum. Molecular Ecology 17:2755–2768.
Mascheretti, S., P. J. P. Croucher, M. Kozanitas, L. Baker, and M. Garbelotto. 2009. Genetic epide-
miology of the sudden oak death pathogen Phytophthora ramorum in California. Molecular
Ecology 18:4577–4590.
Meentemeyer, R. K., D. M. Rizzo, W. Mark, and E. Lotz. 2004. Mapping the risk of establishment
and spread of sudden oak death in California. Forest Ecology and Management 200:195–214.
Meentemeyer, R. K., B. L. Anacker, W. Mark, and D. M Rizzo. 2008a. Early detection of emerging
forest disease using dispersal estimation and ecological niche modeling. Ecological Applica-
tions 18:377–390.
Meentemeyer, R. K., N. E. Rank, B. L. Anacker, D. M. Rizzo, and J. H. Cushman. 2008b. Influence
of land-cover change on the spread of an invasive forest pathogen. Ecological Applications
18:159–171.
Meentemeyer, R. K., N. E. Rank, D. A. Shoemaker, C. B. Oneal, A. C. Wickland, K. Frangioso, and
D. M. Rizzo. 2008c. Impact of sudden oak death on tree mortality in the Big Sur ecoregion of
California. Biological Invasions 10:1243–1255.
Meentemeyer, R. K., N. J. Cunniffe, A. R. Cook, R. D. Hunter, D. M. Rizzo, and C. A. Gilligan. 2011.
Application of stochastic epidemiological models to realistic landscapes: Spread of the sudden
oak death pathogen in California (1990–2030). Ecosphere 2:Article 17.
Monahan, W. B., J. Tse, W. D. Koenig, and M. Garbelotto. 2008. Preserved specimens suggest
non-native origins of three species of Phytophthora in California. Mycological Research 112:
757–758.
Oak, S. W., J. Hwang, S. N. Jeffers, and B. M. Tkacz. 2010. Phytophthora ramorum in USA streams
from the national early detection survey of forests. In Proceedings of the sudden oak death
fourth science symposium, edited by S. J. Frankel, J. T. Kliejunas, and K. M. Palmieri. Albany,
CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Station. Gen.
Tech. Rep. PSW-GTR-229. Pp. 353–354.
APPENDIX A 323
Osterbauer, N. K., J. A. Griesbach, and J. Hedberg. 2004. Surveying for and eradicating Phy-
tophthora ramorum in agricultural commodities. Plant Health Progress, doi:10.1094/
PHP-2004-0309-02-RS.
Parke, J. L., N. J. Grünwald, C. Lewis, and V. Fieland. 2010a. A systems approach for detecting
sources of Phytophthora contamination in nurseries. In Proceedings of the sudden oak death
fourth science symposium, edited by S. J. Frankel, J. T. Kliejunas, and K. M. Palmieri. Albany,
CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Station. Gen.
Tech. Rep. PSW-GTR-229. Pp. 67–68.
Parke, J. L., J. Pscheidt, R. Regan, J. Hedberg, and N. J. Grünwald. 2010b. The Phytophthora online
course: Training for nursery growers. In Proceedings of the sudden oak death fourth science
symposium, edited by S. J. Frankel, J. T. Kliejunas, and K. M. Palmieri. Albany, CA: U.S. De-
partment of Agriculture, Forest Service, Pacific Southwest Research Station. Gen. Tech. Rep.
PSW-GTR-229. Pp. 355.
Prospero, S., N. J. Grunwald, L. M. Winton, and E. M. Hansen. 2009. Migration patterns of the emerg-
ing plant pathogen Phytophthora ramorum on the west coast of the United States of America.
Phytopathology 99:739–749.
Rizzo, D. M. 2005. Exotic species and fungi: Interactions with fungal, plant and animal communi-
ties, In The fungal community (3rd ed.), edited by J. Dighton, P. Oudemans, and J. White. Boca
Raton, FL: CRC Press. Pp. 857–877.
Rizzo, D. M., and M. Garbelotto. 2003. Sudden oak death: Endangering California and Oregon forest
ecosystems. Frontiers in Ecology and the Environment 1:197–204.
Rizzo, D. M., M. Garbelotto, J. M. Davidson, G. W. Slaughter, and S. Koike. 2002. Phytophthora
ramorum as the cause of extensive mortality of Quercus spp. and Lithocarpus densiflorus in
California. Plant Disease 86:205–214.
Rizzo, D. M., M. Garbelotto, and E. M. Hansen. 2005. Phytophthora ramorum: Integrative research
and management of an emerging pathogen in California and Oregon forests. Annual Review of
Phytopathology 43:309–335.
Spaulding, H. L., and L. K. Rieske. 2011. A glimpse at future forests: Predicting the effects of Phy-
tophthora ramorum on oak forests of southern Appalachia. Biological Invasions 6:1367–1375.
Swiecki, T. J., and E. Bernhardt. 2008. Increasing distance from California bay reduces the risk and
severity of Phytophthora ramorum canker in coast live oak. In Proceedings of the third sudden
oak death fourth science symposium, edited by S. J. Frankel, J. T. Kliejunas, and K. M. Palmieri.
Albany, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Sta-
tion. Gen. Tech. Rep. PSW-GTR-214. Pp. 181–194.
Vaclavik, T., A. Kanaskie, E. M. Hansen, J. L. Ohmann, and R. K. Meentemeyer. 2010. Predicting
potential and actual distribution of sudden oak death in Oregon: Prioritizing landscape contexts
for early detection and eradication of disease outbreaks. Forest Ecology and Management
260:1026–1035.
Valachovic, Y., C. Lee, J. Marshall, and H. Scanlon. 2008. Wildland management of Phytophthora
ramorum in northern California forests. In Proceedings of the third sudden oak death fourth
science symposium, edited by S. J. Frankel, J. T. Kliejunas, and K. M. Palmieri. Albany, CA:
U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Station. Gen. Tech.
Rep. PSW-GTR-214. Pp. 305–312.
———. 2010. Forest treatment strategies for Phytophthora ramorum. In Proceedings of the sudden
oak death fourth science symposium, edited by S. J. Frankel, J. T. Kliejunas, and K. M. Palmieri.
Albany, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Sta-
tion. Gen. Tech. Rep. PSW-GTR-229. Pp. 239–248.
Venette, R. C., and S. D. Cohen. 2006. Potential climatic suitability for establishment of Phytophthora
ramorum within the contiguous United States. Forest Ecology and Management 231:18–26.
324 FUNGAL DISEASES
Vettraino, A. M., S. Sukno, A. Vannini, and M. Garbelotto. 2010. Diagnostic sensitivity and specificity
of different methods used by two laboratories for the detection of Phytophthora ramorum on
multiple natural hosts. Plant Pathology 59:289–300.
Werres, S., R. Marwitz, W. A. Man in’t Veld, A. W. A. M. De Cock, P. J. M. Bonants, M. De Weerdt,
K. Themann, E. Ilieva, and R. P. Baayen. 2001. Phytophthora ramorum sp. nov., a new pathogen
on Rhododendron and Viburnum. Mycological Research 105:1155–1165.
Yakabe, L. E., C. L. Blomquist, S. L. Thomas, and J. D. MacDonald. 2009. Identification and fre-
quency of Phytophthora species associated with foliar diseases in California ornamental nurser-
ies. Plant Disease 93:883–890.
A17
Abstract
Empirical evidence from multiple sources shows the Earth has been warm-
ing since the late 19th century. More recently, evidence for this warming trend
is strongly supported by satellite data since the late 1970s from the cryosphere,
atmosphere, oceans, and land. Those data confirm increasing temperature trends
and their consequences (e.g., reduced Arctic sea ice, rising sea level, ice sheet
mass loss, etc.). At the same time, satellite observations of the Sun show remark-
ably stable solar cycles since the late 1970s, when direct observations of the
Sun’s total solar irradiance began. Numerical simulation models, driven in part
by assimilated satellite data, suggest that future warming trends will lead to not
only a warmer planet, but also a wetter and drier climate, depending on location,
in a fashion consistent with large-scale atmospheric processes. Continued global
warming poses new opportunities for the emergence and spread of fungal disease,
as climate systems change at regional and global scales, and as animal and plant
species move into new niches.
Our contribution to this proceedings is organized as follows: First, we re-
view empirical evidence for a warming Earth. Second, we show the Sun is not
responsible for the observed warming. Third, we review numerical simulation
50╛╛Laboratory for Hydrospheric and Biospheric Science, NASA/Goddard Space Flight Center,
Greenbelt, MD 20771.
51╛╛Oak Ridge Associated Universities (ORAU).
52╛╛GESTAR–Universities Space Research Association (USRA), Columbia, MD 21044.
53╛╛USDA/ARS Center for Medical, Agricultural, and Veterinary Entomology, Gainesville, FL.
APPENDIX A 325
modeling results that project these trends into the future, describing the pro-
jected abiotic environment of our planet in the next 40 to 50 years. Fourth, we
illustrate how Rift Valley fever outbreaks have been linked to climate, enabling
a better understanding of the dynamics of these diseases, and how this has led
to the development of an operational predictive outbreak model for this disease
in Africa. Fifth, we project how this experience may be applicable to predicting
outbreaks of fungal pathogens in a warming world. Last, we describe an example
of changing species ranges due to climate change, resulting from recent warming
in the Andes and associated glacier melt that has enabled amphibians to colonize
higher elevation lakes, only to be followed shortly by the emergence of fungal
disease in the new habitats.
FIGURE A17-1╇ Summary of observations that show the Earth is warming (red arrows)
Figure A17-1.eps
while the Sun has been constant over the same period of time.
bitmap
Surface Thermometers
Four global surface temperature datasets are available from the following
locations: (1) National Aeronautics and Space Administration (NASA)/Goddard
Institute of Space Studies (Hansen et al., 2010), (2) National Oceanic and Atmo-
spheric Administration’s National Climatic Data Center (Smith et al., 2008), (3)
the University of East Anglia’s Climate Research Unit (Rayner et al.. 2006), and
(4) the Japanese Meteorological Agency.54 These four datasets all use the same
input data and differ only in interpolation techniques between sparse observa-
tions, how the polar regions are treated, and the reference period for which means
are calculated. Not surprisingly, they are very similar (Figure A17-2).
54╛╛See http://www.jma.go.jp/jma/indexe.html.
APPENDIX A 327
FIGURE A17-2╇ A comparison of the existing four global surface temperature datasets
that are used in climate analyses.Figure A17-2.eps
These datasets are based on the same input data and dif-
fer by interpolation among stations, treatment of missing data, and the length of the record.
type is outlined
The data in this figure have been adjusted to a common baseline period.
SOURCE: Figure courtesy of Robert Simmon, NASA Earth Observatory, provided by
Compton J. Tucker NASA Goddard Space Flight Center.
Sea Level
Sea level is of direct interest to climate science because it varies directly with
global mean temperature over short time scales. Temperature affects sea level
through two mechanisms: (1) sea level rises through the thermal expansion of
water as it warms, or it falls through thermal contraction of water as it cools; and
(2) warmer global temperatures melt ice stored on land in glaciers and ice sheets,
and the resulting ice loss raises sea level, while cooler global temperatures result
in more water being stored on land in glaciers and ice sheets and sea level falls.
Thus sea-level variations are an excellent, unambiguous indicator of planetary
cooling or warming. For example, at the last glacial maximum, occurring before
20,000 years ago, sea level was >100 m lower than it is now (Clark et al., 2009).
This huge quantity of water was stored on land in the form of glaciers and ice
sheets (Lambeck et al., 2010).
Although tide gauges provide centennial-scale sea-level records from nearly
10 locations around the world, these few locations are insufficient for a global
study of sea level. Researchers have also measured vertical accretion rates in
salt marshes as a sea-level proxy, using radiocarbon, pollen, foraminifera, and
other markers (reviewed in Mitchum et al., 2010). Since 1993, however, radar
altimeters have measured sea level globally and directly with a high precision
(Figure A17-3).
We have briefly reviewed global surface thermometer data, atmospheric
temperature profile data from satellites, variations in Arctic sea ice, and sea-
level data. All of these data unambiguously show the effects of increasing global
temperatures.
FIGURE A17-3╇ Sea-level rise based on radar altimeters from TOPEX and Jason, with
seasonal variations removed. Figure A17-3.eps
SOURCE: Mitchum et al. (2010). bitmap
sure. Each grid cell interacts with adjacent cells horizontally and vertically to
simulate climate (Figure A17-5). Model interactions are governed by systems
of differential equations and incorporate climate-forcing factors such as land
cover change, volcanic aerosols, and increasing greenhouse gas concentrations.
Weather and climate models have been shown to be realistic at reproducing the
global temperature and precipitation patterns of the 20th century and are widely
used in weather and climate research (Delworth et al., 2006).
Climate model simulations, incorporating increasing greenhouse gas concen-
trations in the atmosphere, have been used to extrapolate precipitation patterns
into the 21st century as surface temperatures increase. Several of these climate
model simulation predictions can be described as “the wet getting wetter and the
dry getting drier” (Held and Soden, 2006). The displacement of arid and semi-
arid zones northward results from an expansion from the Hadley circulation cell
under global warming (Figure A17-6) (Lu et al., 2007). These changes in climate
have direct impacts on vegetation and biodiversity across the globe, including
species range shifts, changing phenology, new invasive species, and new disease
outbreaks (Parmesan and Yohe, 2003; Walther et al., 2002).
330 FUNGAL DISEASES
FIGURE A17-4╇ A comparison between the total solar irradiance (top) and the NASA/
GISS surface temperature data (bottom),
Figure both from 1979 to 2010. This shows the sun is
A17-4.eps
not responsible for the 1979 to 2010 increased surface temperatures.
bitmap
FIGURE A17-5╇ Representation of a general circulation model illustrating the grid cell
nature of the model on the right,Figure
while on A17-5.eps
the left, many of the different important com-
ponents of these models are shown. bitmap
SOURCE: Figure courtesy of the Center for Multiscale Modeling of Atmospheric Pro-
cesses (CMMAP), Colorado State University, http://www.cmmap.org/learn/modeling/
whatIs2.html.
FIGURE A17-6╇ Change in precipitation between the 1971–2000 average and the 2091–
2100 average in inches of liquidFigure
water/yearA17-6.eps
(Held and Soden, 2006).
SOURCE: Geophysical Fluid Dynamics Laboratory, National Oceanic and Atmospheric
Administration.
bitmap
332 FUNGAL DISEASES
records from numerous stations in Kenya between 1950 and 1982, it was deter-
mined that periods with extended positive surplus rainfall corresponded to periods
when Rift Valley fever epizootics occurred. Widespread, frequent, and persistent
rainfall was shown to be a prominent feature of all epizootic periods. Heavy rain-
fall raises the level of the water table in certain areas, flooding grassland depres-
sions that are the habitat of the immature stages of certain ground-pool–breeding
mosquitoes of the genus Aedes. These findings have been collaborated by find-
ings in southern Africa (Swanepoel, 1976) and West Africa (Bicout and Sabatier,
2004). Rift Valley fever virus is thought to be initially transmitted transovarially
in these species. Under prolonged flooded conditions, large numbers of Culex
species mosquitoes emerge and are an amplification vector for Rift Valley fever.
Following the development of these conditions, Rift Valley fever first occurs in
animals and subsequently in humans.
Linthicum et al. (1999) established that outbreaks of Rift Valley fever are
closely coupled with above normal rainfall that is associated with the occur-
rence of the warm phase of El Niño/Southern Oscillation (ENSO) (Cane, 1983;
Nicholson, 1986; Ropelewski and Halpert, 1987) and warm events in the equato-
rial western Indian Ocean (Anyamba et al., 2002; Birkett et al., 1999; Saji et al.,
1999). Such warm ocean events precede by 2 to 3 months above normal and
extended rainfall over East Africa, and are further enhanced when both the sea
surface anomalies in the western Indian Ocean and equatorial central-eastern
Pacific are synchronized. More than 90 percent of Rift Valley fever outbreak
events since 1950 have occurred during warm ENSO events (Linthicum et al.,
1999) (Figure A17-7). The interepizootic period is dominated by La Niña events
(the cold phase of ENSO), which results in drought in East Africa and wetter
than normal conditions in southern Africa (Anyamba et al., 2002; Nicholson
and Entekhabi, 1986). Recent evidence shows that Rift Valley fever outbreaks
in southern Africa are coupled with La Niña patterns (Anyamba et al., 2010).
Interannual variability, in part driven by ENSO events with differential impacts
on rainfall anomaly patterns in Eastern and Southern Africa, largely influences
the temporal outbreak patterns of Rift Valley fever.
Our work on Rift Valley fever prediction thus uses climate data to inform
us when and where regionally we should expect outbreaks. Subsequent detailed
daily satellite observations identify where outbreaks will occur with a high degree
of geographical specificity (~60 percent).
4 1951-3 1957 1961-3 1968-9 1977-9 1982-3 1989 1997-8 2006 2010
-1
-2
-3
-5
1951 1955 1959 1963 1967 1971 1975 1979 1983 1987 1991 1995 1999 2003 2007 2011
FIGURE A17-7╇ Rift Valley fever major outbreak events plotted against time and the Southern Oscillation Index, a measure of the phase of El
Niño/Southern Oscillation events. Most Rift Valley fever outbreak events have occurred during the warm phase of ENSO (negative Southern
Oscillation Index shown in blue).
SOURCE: Linthicum et al. (1999), and updated.
Figure A17-7.eps
333
landscape
type is 3.69 point--some type is outlined
334 FUNGAL DISEASES
A B
C D
FIGURE A17-8╇ Summary Rift Valley fever (RVF) risk maps for (A) Eastern Africa:
September 2006–May 2007; (B)Figure A17-8.eps
Sudan: May 2007–December 2007; (C) Southern Africa:
4 bitmaps w vector
September 2007–May 2008; and (D) Madagascar: type & ruling
September 2007–May 2008. Areas
shown in green represent Rift Valley fever potential epizootic areas. Areas shown in red
represent pixels that were mapped by the prediction system to be at risk for RVF activity
during the respective time periods. Blue dots indicate human cases identified to be with
the Rift Valley fever risk areas, while yellow dots represent human cases in areas not
mapped to be at risk.
SOURCE: Adapted from Anyamba et al. (2010).
336 FUNGAL DISEASES
FIGURE A17-10╇ False-color Landsat satellite data (RGB 642) showing glaciers as the
Figure A17-10.eps
blue colors. The green colors represent green vegetation and the red colors represent areas
of rock, sand, and soil. bitmap
SOURCE: Figure provided by Karina Yager, NASA-Goddard Space Flight Center.
Conclusions
The Earth’s climate is warming and our Sun is not responsible. Weather and
climate simulation models project even warmer temperatures by the middle of
this century, with some areas getting wetter and others drier. These changing pat-
terns of temperature and precipitation will alter endemic areas for various plant
and animal diseases, including fungal pathogens. We reviewed how knowledge
of climatic linkages is being used to predict the outbreak regions of Rift Valley
fever in Africa, complemented by detailed satellite observations to identify spe-
cific locales where control measures should be undertaken. We advocate a similar
approach to identify areas where fungal diseases may emerge: understand the
biology of specific fungal pathogens; use satellite data to establish temperature
and precipitation climatology in the areas of interest; associate this information
with documented fungal outbreaks; and use this knowledge in conjunction with
satellite data to predict the impacts of a changing and variable climate on fungal
pathogens.
References
Anyamba, A., K. J. Linthicum, R. Mahoney, C. J. Tucker, and P. W. Kelley. 2002. Mapping potential
risk of Rift Valley fever outbreaks in African savannas using vegetation index time series data.
Photogrammatic Engineering & Remote Sensing 68(2):137–145.
APPENDIX A 339
Anyamba, A., J. Chretien, J. Small, C. J. Tucker, and K. J. Linthicum. 2006. Developing global
climate anomalies suggest potential disease risks for 2006–2007. International Journal of
Health Geographics 5:60. http://www.ij-healthgeographics.com/content/5/1/60 (accessed June
14, 2011).
Anyamba, A., J.-P. Chretien, J. Small, C. J. Tucker, P. Formenty, J. H. Richardson, S. C. Britch,
D.€C. Schnabel, R. L. Erickson, and K. J. Linthicum. 2009. Prediction of the Rift Valley fever
outbreak in the Horn of Africa 2006–2007. Proceedings of the National Academy of Sciences,
USA 106(3):955–959.
Anyamba, A., K. J. Linthicum, J. Small, S. C. Britch, E. Pak, S. de La Rocque, P. Formenty, A. W.
Hightower, R. F. Breiman, J.-P. Chretien, C. J. Tucker, D. Schnabel, R. Sang, K. Haagsma, M.
Latham, H. B. Lewandowski, S. O. Magdi, M. A. Mohamed, P.M. Nguku, J.-M. Reynes, and R.
Swanepoel. 2010. Prediction, assessment of the Rift Valley fever activity in East and Southern
Africa 2006–2008 and possible vector control strategies. American Journal of Tropical Medicine
and Hygiene 81(5):43–51.
Baylis, M., P. Mellor, and R. Meiswinkel. 1999. Horse sickness and ENSO in South Africa. Nature
397(6720):574.
Bicout, D. J., and P. Sabatier. 2004. Mapping Rift Valley fever vectors and prevalence using rainfall
variations. Vector-Borne and Zoonotic Diseases 4(1):33–42.
Birkett, C., R. Murtugudde, and T. Allan. 1999. Indian Ocean climate event brings floods to East
Africa’s lakes and the Sudd Marsh. Geophysical Research Letters 26(8):1031–1034.
Bouma, M., and C. Dye. 1997. Cycles of malaria associated with El Niño in Venezuela. Journal of
the American Medical Association 278(21):1772–1774.
Bouma, M., C. Dye, and J. van der Kay. 1996. Falciparum malaria and climate change in the
Northwest Frontier Province of Pakistan. American Journal of Tropical Medicine and Hygiene
55(2):131–137.
Bradley, R. S., M. Vuille, H. F. Diaz, and W. Vergara. 2006. Threats to water supplies in the tropical
Andes. Science 312(5781):1755–1756.
Cane, M. A. 1983. Oceanographic events during El Niño. Science 222(4629):77–90.
Checkley, W., L. D. Epstein, R. H. Gilman, D. Figuerora, R. I. Cama, and J. A. Patz. 2000. Effects
of El Niño and ambient temperature on hospital admissions for diarrheal diseases in Peruvian
children. Lancet 355(9202):442–450.
Chretien, J.-P., A. Anyamba, S. A. Bedno, F. Breiman, R. Sang, K. Sergon, A. M. Powers, C. O.
Onyango, J. Small, C. J. Tucker, and K. J. Linthicum. 2007. Drought-associated chikungunya
emergence along coastal East Africa. American Journal of Tropical Medicine and Hygiene
76(3):405–407.
Clark, P. U., A. S. Dyke, J. D. Shakun, A. E. Carlson, J. Clark, B. Wohlfarth, J. X. Mitrovica, S. W.
Hostetler, and A. M. McCabe. 2009. The last glacial maximum. Science 325(5941):710–714.
Colwell, R. R. 1996. Global climate and infectious disease: The cholera paradigm. Science 274(5295):
2025–2031.
Daszak, P., L. Berger, A. A. Cunningham, A. D. Hyatt, D. E. Green, and R. Speare. 1999. Emerging
infectious diseases and amphibian population declines. Emerging Infectious Diseases 5(6):
735–748.
Davies, F. G., K. J. Linthicum, and A. D. James. 1985. Rainfall and epizootic Rift Valley fever. Bul-
letin of the World Health Organization 63(5):941–943.
Delworth, T. L., A. J. Broccoli, A. Rosati, R. J. Stouffer, V. Balaji, J. A. Beesley, W. F. Cooke, K. W.
Dixon, J. Dunne, K. A. Dunne, J. W. Durachta, K. L. Findell, P. Ginoux, A. Gnanadesikan, C. T.
Gordon, S. M. Griffies, R. Gudgel, M. J. Harrison, I. M. Held, R. S. Hemler, L. W. Horowitz,
S. A. Klein, T. R. Knutson, P. J. Kushner, A. R. Langenhorst, H. C. Lee, S. J. Lin, J. Lu, S. L.
Malyshev, P. C. D. Milly, V. Ramaswamy, J. Russell, M. D. Schwarzkopf, E. Shevliakova, J. J.
Sirutis, M. J. Spelman, W. F. Stern, M. Winton, A. T. Wittenberg, B. Wyman, F. Zeng, and R.
Zhang. 2006. GFDL’s CM2 global coupled climate models—Part 1: Formulation and simulation
characteristics. Journal of Climate 19(5):643–674.
340 FUNGAL DISEASES
Diaz, H. F., and N. E. Graham. 1996. Recent changes in tropical freezing heights and the role of sea
surface temperature. Nature 383(6596):152–155.
Ebi, K. L., K. A. Exuzides, E. Lau, M. Kelsh, and A. Barnston. 2001. Association of normal weather
periods and El Nino events with hospitalization for viral pneumonia in females: California,
1983–1998. American Journal of Public Health 91(8):1200–1208.
Glass, G. E., T. L. Yates, J. B. Fine, T. M. Shields, J. B. Kendall, A. G. Hope, C. A. Parmenter, C. J.
Peters, T. G. Ksiazek, C.-S. Li, J. A. Patz, and J. N. Mills. 2002. Satellite imagery character-
izes local animal reservoir populations of Sin Nombre virus in the southwestern United States.
Proceedings of the National Academy of Sciences, USA 99(26):16817–16822.
Halloy, S., A. Seimon, and K. Yager. 2006. Multidimensional (climate, biodiversity, socioeconom-
ics, agriculture) context of changes in land use in the Vilcanota watershed, Peru. In Land use
change and mountain biodiversity, edited by E. Spehn, M. Liberman-Cruz, and C. Körner. Ft.
Lauderdale, FL: CRC Press. Pp. 323–337.
Hansen, J., R. Ruedy, M. Sato, and K. Lo. 2010. Global surface temperature change. Reviews of
Geophysics 48(RG4004, 29):1–29.
Held, I. M., and B. J. Soden. 2006. Robust responses of the hydrological cycle to global warming.
Journal of Climate 19(21):5686–5699.
Hopp, M. J., and J. A. Foley. 2003. Worldwide fluctuations in dengue fever cases related to climate
variability. Climatic Research 25(1):85–94.
Kovats, R. S., S. J. Edwards, S. Hajat, B. G. Armstrong, K. L. Ebi, B. Menne, and the Collaborating
Group. 2004. The effect of temperature on food poisoning: A time-series analysis of salmonel-
losis in ten European countries. Epidemiological Infections 132(3):443–453.
Kwok, R. 2002. Sea ice concentration estimates from satellite passive microwave radiometry and
openings from SAR ice motion. Geophysical Research Letters 29(9):25-1–25-4.
Lambeck, K., C. D. Woodroffe, F. Antonioli, M. Anzidei, W. R. Gehrels, J. Laborel, and A. J. Wright.
2010. Paleoenvironmental records, geophysical modeling, and reconstructions of sea-level
trends and variability on centennial and longer timescales. In Understanding sea level rise and
variability, edited by P. L. Woodworth, J. A. Church, T. Aarup, and W. S. Wilson. New York:
Wiley-Blackwell. Pp. 61–121.
Linthicum, K. J., A. Anyamba, C. J. Tucker, P. W. Kelley, M. F. Myers, and C. J. Peters. 1999.
Climate and satellite indicators to forecast Rift Valley fever epidemics in Kenya. Science
285(5426):397–400.
Linthicum, K. J., S. C. Britch, A. Anyamba, J. Small, C. J. Tucker, J.-P. Chretien, and R. Sithipraasasna.
2008. Ecology of disease: The intersection of human and animal health. In Vector-borne dis-
eases—Understanding the environmental, human health, and ecological considerations, Insti-
tute of Medicine of The National Academies. Washington, DC: National Academy Press. Pp.
78–88.
Lockwood, M. 2009. Solar change and climate: An update in the light of the current exceptional solar
minimum. Proceedings of the Royal Society, Series A 466(2114):303–329.
Lockwood, M., and C. Frohlich. 2007. Recent oppositely-directed trends in solar climate forc-
ings and the global mean surface air temperature. Proceedings of the Royal Society, Series A
463(2086):2447–2460.
———. 2008. Recent oppositely-directed trends in solar climate forcings and the global mean surface
air temperature. Different reconstructions of the total solar irradiance variation and dependence
on response time scale. Proceedings of the Royal Society, Series A. 464(2094):1367–1385.
Lu, J., G. A. Vecchi, and T. Reichler. 2007. Expansion of the Hadley cell under global warming.
Geophysical Research Letters 34(L06805):1–5.
Lyman, J. M., S. A. Good, V. V. Gouretski, M. Ishii, G. C. Johnson, M. D. Palmer, S. M. Smith, and
J. K. Willis. 2010. Robust warming of the global upper ocean. Nature 465(7296):334–337.
Mitchum, G. T., R. S. Nerem, M. A. Merrifield, and W. R. Gehrels. 2010. Modern sea-level-change
estimates. In Understanding sea level rise and variability, edited by P. L. Woodworth, J. A.
Church, T. Aarup, and W. S. Wilson. New York: Wiley-Blackwell. Pp. 122–142.
APPENDIX A 341
A18
Abstract
Amphibian biodiversity is currently facing a severe crisis having recently
experienced declines in 42 percent of all species, and as many as 32 percent are
56╛╛Department of Biology, San Francisco State University, San Francisco, CA.
57╛╛Department of Ecology, Evolution and Marine Biology, University of California Santa Barbara,
Santa Barbara, CA.
58╛╛Department of Biology, James Madison University, Harrisonburg, VA.
APPENDIX A 343
threatened with imminent extinction. The most alarming extinctions and declines
have occurred enigmatically in protected, apparently pristine habitats. An emerg-
ing infectious disease, chytridiomycosis, is directly linked to the recent extinction
or serious decline of hundreds of amphibian species and is increasingly proposed
as a primary threat to amphibians. Chytridiomycosis, caused by the fungal patho-
gen Batrachochytrium dendrobatidis (Bd), infects the skin of amphibians and has
been described as causing the greatest loss of vertebrate biodiversity due to dis-
ease in recorded history. The severity of the current amphibian biodiversity crisis
suggests that Bd is a fundamentally new challenge to amphibians from previous
global and environmental changes. While many amphibian species are susceptible
to this fungal pathogen, others are silent carriers exhibiting no signs of disease.
How amphibian hosts survive with Bd infection is still unknown; however, host
immunity, differences in pathogen virulence, and environmental differences that
may limit growth and reproduction of the pathogen have all been proposed as pos-
sible mechanisms that could lead to Bd-infected host survival. Here we examine
amphibian declines in one of the best-documented systems, the Sierra Nevada of
California, and review the role that the amphibian skin microbiome may play in
host–pathogen dynamics of the chytridiomycosis-amphibian host system.
Introduction
The amphibians are long-term survivors (existing on earth for more than
350 million years) that endured four previous mass extinctions (e.g., 95 percent
of all living species were lost in the Permian-Triassic extinction) (Wake and
Vredenburg, 2008). Through these extinctions, not only did all three orders of
amphibians escape extinction, but most families and genera survived (Wake and
Vredenburg, 2008). Today, the amphibians, presently including more than 6,800
species (AmphibiaWeb, 2011), are the most threatened group of vertebrates with
over 40 percent of species in decline and over 30 percent threatened with extinc-
tion (Stuart et al., 2004). There are many potential causes for the declines, but
an emerging infectious disease, chytridiomycosis (Longcore et al., 1999), caused
by the infectious fungal pathogen Batrachochytrium dendrobatidis (Bd) is the
most alarming (Daszak et al., 2000). This fungal pathogen is associated with
the recent decline or extinction of more than 200 species of amphibians (Fisher
et al., 2009; Skerratt et al., 2007). Epizootic waves of chytridiomycosis have
been identified in Australia, Central and South America, and the western United
States (Fisher et al., 2009). In each case the epizootic caused mass mortality and
collapse of amphibian faunas (Berger et al., 1998, Lips et al., 2006, Vredenburg
et al., 2010a). Once Bd invades naive amphibian host populations, they can col-
lapse very quickly. In Panamanian sites, 50 percent of species were extirpated 4
to 6 months after Bd invaded (Lips et al., 2006). The surviving species declined
in population size by 80 percent (Lips et al., 2006). In California, a moving epi-
zootic wave of Bd is causing the collapse of entire metapopulations of ranid frogs
344 FUNGAL DISEASES
in some of the most protected and pristine areas in the United States (Vredenburg
et al., 2010a). In Australia, most of the damage was done before the pathogen
was identified (Laurance et al., 1996), but post hoc studies estimate that Bd was
responsible for the decline and disappearance of a majority of amphibian species
along the diverse eastern tropical montane areas (Fisher et al., 2009).
Exnct
Extant (1997) B
FIGURE A18-1╇ Decline of (A) Sierra Nevada mountain yellow-legged frog, Rana sier-
rae, and (B) southern mountain yellow-legged frog, Rana muscosa, in California, USA.
Yellow points indicate extant populations in 1997 and red points indicate extinct popula-
tions compared to historically documented sites.
SOURCE: Adapted from Vredenburg et al. (2007).
A18-1
non-native trout, which have contributed significantly to the decline of the two
species of yellow-legged frog (Knapp and Matthews, 2000; Vredenburg, 2004).
In 2004 a whole-lake experimental study determined that predation on tadpoles
by introduced trout was causing frog population extirpations (Vredenburg, 2004).
The same study also demonstrated that threatened frog populations would recover
quickly when habitat was restored by removing the non-native fish (Vredenburg,
2004). Additional research showed that habitat restoration (to the fishless con-
dition) worked across the mountain range for both yellow-legged frog species
(Knapp et al., 2007). Unfortunately, further research ultimately revealed that
amphibian declines in the Sierra Nevada, perhaps like other areas around the
world, are complex in nature and can be caused by multiple factors that may act
346 FUNGAL DISEASES
populations of frogs and that infection prevalence quickly reached 100 percent
(Figure A18-3, panel B), but mass mortality and host population collapse did not
occur until the average Bd infection load (or intensity of infection on individu-
als) for a population reached more than 104 zoospore equivalents (Figure A18-3,
panels A and C; Vredenburg et al., 2010a). This condition was later named the
“Vredenburg 10,000 Zoospore Rule” (Kinney et al., 2011) and was shown to also
predict mortality in neotropical salamanders (Cheng et al., 2011).
In the Sierra Nevada, infection by Bd did not always lead to epizootic events
and host population collapse. About 100 km north of our SEKI study area in
and around Yosemite National Park, we describe the Bd–frog dynamics under
stable enzootic conditions (Briggs et al., 2010), which contrast sharply with
conditions we describe above in the same species of yellow-legged frog in SEKI
(Vredenburg et al., 2010a). We discovered that in the Yosemite area, frog host
population dynamics are strikingly different and that host populations survive, at
least for a decade, despite infection by Bd (Briggs et al., 2010). We determined
that while infection prevalence (proportion of infected hosts in a population)
could be high (more than 60 percent), the infection intensity of hosts remained
low (Briggs et al., 2010), well below the Vredenburg 10,000 Zoospore Rule
(Kinney et al., 2011). In fact, several other studies have found that host amphib-
ians do not die when Bd infection intensities remain low (Cheng et al., 2011;
Kinney et al., 2011; Kriger et al., 2007; Retallick et al., 2004). Why the dynamics
of the host–pathogen interaction is so different in different sites and among differ-
ent species remains a mystery, but we believe the key to host survival is the low
infection intensities on individual host frogs (Briggs et al., 2010; Kinney et al.,
2011; Vredenburg et al., 2010a) and salamanders (Cheng et al., 2011).
Figure A18-3.eps
FIGURE A18-3╇Frog–Bd dynamics in eight intensively sampled populations in Mile-
bitmap
stone and Sixty Lake basins before and after detection of Bd: (A) frog counts (adults +
subadults) from visual encounter surveys; (B) infection prevalence, defined as the fraction
of skin swabs collected from each population on each date positive for Bd; and (C) infec-
tion intensity, defined as the average zoospore equivalents on swabs collected from each
population on each date. Data are from frog populations that were sampled more than
once per year, that experienced more than 80 percent declines by the end of 2006, and for
which the decline in the number of frogs was greater than 10. This last criterion excluded
populations that were very small before Bd arrival. Populations are aligned along the x axis
such that “0” represents the date on which each frog population began to decline. This was
calculated for each population by determining the date at which the number of postmeta-
morphic frogs dropped below 20 percent of the average population count before that point.
SOURCE: Vredenburg et al. (2010a).
350 FUNGAL DISEASES
and Trueb, 1986). Many species are well known for this characteristic, such as the
poison dart frogs of the neotropics (genus Dendrobates) that produce neurotoxins
strong enough to incapacitate monkeys hit by blow-darts from hunters who rub
the darts on the skin of the frogs, or the toxic newts (genus Taricha) that produce
the same chemical defense compounds, known as tetrodotoxins, found in deadly
puffer fish.
Amphibians, although generally tied to water, have colonized many diverse
habitats, including many with no standing or running water (Duellman, 1999). In
fact, many diverse species have developed reproductive modes that free them in
some capacity or completely from water (e.g., direct developers). The four-toed
salamander (Hemidactylium scutatum), although not a direct developer, is of
interest for this chapter because the female will lay her eggs outside water along
steep banks or edges of ponds and she will brood them or guard them until they
hatch as larvae. Originally it was presumed that the female was guarding her
eggs against egg predators, but behavioral experiments showed that unattended
eggs were not eaten by egg predators; in fact, the eggs were unpalatable to many
predators (Hess and Harris, 2000), and died instead from fungal infections.
Egg-brooding females protect their eggs by inoculating them with colonies of
beneficial bacteria (Harris et al., 2006). A laboratory study later showed that the
bacterium, Janthinobacteria lividum, cultured from female four-toed salamanders
inhibited the growth of Bd (Harris et al., 2006).
Could the information from egg-brooding salamanders in the eastern United
States provide insight into why some populations of yellow-legged frogs in Cali-
fornia survived Bd infection while others did not? In 2005 we collected bacterial
cultures from susceptible and nonsusceptible populations of Rana muscosa and
R. sierrae. We found that individual frogs from nonsusceptible populations were
more likely to contain culturable populations of the symbiotic bacterium Jan-
thinobacteria lividum, the same species of bacteria discovered on the four-toed
salamander (Woodhams et al., 2007). A laboratory study to test whether inocula-
tions of J. lividum could protect susceptible R. muscosa from Bd infections was
then initiated. Adult R. muscosa were collected from the wild in SEKI in an area
ahead of the Bd epizootic wave and were transported to the Harris Laboratory at
James Madison University. The laboratory experiment consisted of three treat-
ments of frogs: one group received J. lividum only, another group received Bd
only, and a third was first inoculated with J. lividum and was then infected with
Bd. As expected, all of the frogs in the Bd-only treatment died from chytridiomy-
cosis within several weeks, but the frogs inoculated with J. lividum before being
infected survived, as did the frogs that only received inoculations of J. lividum,
providing evidence that bacterial symbionts could alter the outcome of Bd infec-
tions in susceptible host populations (Harris et al., 2009).
APPENDIX A 351
Acknowledgments
We thank Harold Werner, David Graber, and Danny Boiano of the National
Park Service for their support of our project and to SEKI for providing the proper
research permits. We especially thank Lilia Torres, without whom we would not
have been able to culture Janthinobacterium lividum, and David Daversa, who
collected most of the field data. We thank Anand Varma for taking wonderful
photographs; Celeste Dodge, Cory Singer, Sam McNally, and Lauren Gillespie
for all the hard work; and Roland Knapp and his field crew for logistics, field
work, and planning. We also thank all of the Forum members and the Institute of
Medicine for the invitation to participate and LeighAnne Olsen for her patience
and kindness to the lead author. Our research is supported by National Science
Foundation Grant EF-0723563 as part of the joint National Science Founda-
tion–National Institutes of Health Ecology of Infectious Disease program and by
startup funds provided to V. Vredenburg by the College of Science and Engineer-
ing at San Francisco State University.
References
AmphibiaWeb: Information on amphibian bioloby and conservation [web application]. Berkeley,
California: AmphibiaWeb. Available: http://amphibiaweb.org/ (accessed 2011).
Berger, L., R. Speare, P. Daszak, D. E. Green, A. A. Cunningham, C. L. Goggin, R. Slocombe, M. A.
Ragan, A. D. Hyatt, K. R. McDonald, H. B. Hines, K. R. Lips, G. Marantelli, and H. Parkes.
1998. Chytridiomycosis causes amphibian mortality associated with population declines in the
rain forests of Australia and Central America. Proceedings of the National Academy of Sciences,
USA 95(15):9031–9036.
Berger, L., A. D. Hyatt, R. Speare, and J. E. Longcore. 2005. Life cycle stages of the amphibian
chytrid Batrachochytrium dendrobatidis. Diseases of Aquatic Organisms 68(1):51–63.
Blaustein, A. R., P. D. Hoffman, D. G. Hokit, J. M. Kiesecker, S. C. Walls, and J. B. Hays. 1994. UV
repair and resistance to solar UV-B in amphibian eggs: A link to population declines? Proceed-
ings of the National Academy of Sciences, USA 91(5):1791–1795.
Boyle, D. G., D. B. Boyle, V. Olsen, J. A. T. Morgan, and A. D. Hyatt. 2004. Rapid quantitative
detection of chytridiomycosis (Batrachochytrium dendrobatidis) in amphibian samples using
real-time TaqMan PCR assay. Diseases of Aquatic Organisms 60(2):141–148.
Bradford, D. F., R. A. Knapp, D. W. Sparling, M. S. Nash, K. A. Stanley, N. G. Tallent-Halsell,
L. L. McConnell, and S. M. Simonich. 2011. Pesticide distributions and population declines of
California, USA, alpine frogs, Rana muscosa and Rana sierrae. Environmental Toxicology and
Chemistry 30(3):682–691.
APPENDIX A 353
Briggs, C. J., R. A. Knapp, and V. T. Vredenburg. 2010. Enzootic and epizootic dynamics of the
chytrid fungal pathogen of amphibians. Proceedings of the National Academy of Sciences, USA
107(21):9695–9700.
Brucker, R. M., R. N. Harris, C. R. Schwantes, T. N. Gallaher, D. C. Flaherty, B. A. Lam, and K. P.
C. Minbiole. 2008. Amphibian chemical defense: Antifungal metabolites of the microsymbiont
Janthinobacterium lividum on the salamander Plethodon cinereus. Journal of Chemical Ecol-
ogy 34(11):1422–1429.
Cheng, T. L., S. M. Rovito, D. B. Wake, and V. T. Vredenburg. 2011. Coincident mass extirpation of
neotropical amphibians with the emergence of the infectious fungal pathogen Batrachochytrium
dendrobatidis. Proceedings of the National Academy of Sciences, USA 108(23):9502–9507.
Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2000. Emerging infectious diseases of wildlife—
threats to biodiversity and human health. Science 287(5452):443–449.
Davidson, C. 2004. Declining downwind: Amphibian population declines in California and historical
pesticide use. Ecological Applications 14(6):1892–1902.
Duellman, W. E. 1999. Patterns of distribution of amphibians. Baltimore: The Johns Hopkins Uni-
versity Press.
Duellman, W. E., and L. Trueb. 1986. Biology of amphibians. New York: McGraw-Hill.
Fisher, M. C., T. W. J. Garner, and S. F. Walker. 2009. Global emergence of Batrachochytrium
dendrobatidis and amphibian chytridiomycosis in space, time, and host. Annual Review of
Microbiology 63:291–310.
Grinnell, J., and T. Storer. 1924. Animal life in the Yosemite. Berkeley: University of California Press.
Harris, R. N., T.Y. James, A. Lauer, M. A. Simon, and A. Patel. 2006. The amphibian pathogen
Batrachochytrium dendrobatidis is inhibited by the cutaneous bacteria of amphibian species.
EcoHealth 3:53–56.
Harris, R. N., R. M. Brucker, J. B. Walke, M. H. Becker, C. R. Schwantes, D. C. Flaherty, B. A.
Lam, D. C. Woodhams, C. J. Briggs, V. T. Vredenburg, and K. P. C. Minbiole. 2009. Skin mi-
crobes on frogs prevent morbidity and mortality caused by a lethal skin fungus. ISME Journal
3(7):818–824.
Hess, Z. J., and R. N. Harris. 2000. Eggs of Hemidactylium scutatum (caudata: Plethodontidae) are
unpalatable to insect predators. Copeia 2000:597–600.
Hyatt, A. D., D. G. Boyle, V. Olsen, D. B. Boyle, L. Berger, D. Obendorf, A. Dalton, K. Kriger, M.
Hero, H. Hines, R. Phillott, R. Campbell, G. Marantelli, F. Gleason, and A. Colling. 2007.
Diagnostic assays and sampling protocols for the detection of Batrachochytrium dendrobatidis.
Diseases of Aquatic Organisms 73(3):175–192.
Kinney, V. C., J. L. Heemeyer, A. P. Pessier, and M. J. Lannoo. 2011. Seasonal pattern of Batracho-
chytrium dendrobatidis infection and mortality in Lithobates areolatus: Affirmation of Vreden-
burg’s “10,000 zoospore rule”. PLoS ONE 6(3):e16708.
Knapp, R. A., and K. R. Matthews. 2000. Non-native fish introductions and the decline of the moun-
tain yellow-legged frog from within protected areas. Conservation Biology 14(2):428–438.
Knapp, R. A., D. M. Boiano, and V. T. Vredenburg. 2007. Removal of nonnative fish results in
population expansion of a declining amphibian (mountain yellow-legged frog, Rana muscosa).
Biological Conservation 135(1):11–20.
Kriger, K. M., F. Pereoglou, and J.-M. Hero. 2007. Latitudinal variation in the prevalence and inten-
sity of chytrid (Batrachochytrium dendrobatidis) infection in eastern Australia. Conservation
Biology 21(5):1280–1290.
Lam, B. A., J. B. Walke, V. T. Vredenburg, and R. N. Harris. 2010. Proportion of individuals with
anti-Batrachochytrium dendrobatidis skin bacteria is associated with population persistence in
the frog Rana muscosa. Biological Conservation 143(2):529–531.
Laurance, W. F., K. R. McDonald, and R. Speare. 1996. Epidemic disease and the catastrophic decline
of Australian rain forest frogs. Conservation Biology 10(2):406–413.
354 FUNGAL DISEASES
Lips, K. R., F. Brem, R. Brenes, J. D. Reeve, R. A. Alford, J. Voyles, C. Carey, L. Livo, A. P. Pessier,
and J. P. Collins. 2006. Emerging infectious disease and the loss of biodiversity in a neotropical
amphibian community. Proceedings of the National Academy of Sciences, USA 103:3165–3170.
Longcore, J. E., A. P. Pessier, and D. K. Nichols. 1999. Batrachocytrium dendrobatidis gen. et sp.
nov., a chytrid pathogenic to amphibians. Mycologia 91:219–227.
Morgan, J. A. T., V. T. Vredenburg, L. J. Rachowicz, R. A. Knapp, M. J. Stice, T. Tunstall, R. E.
Bingham, J. M. Parker, J. E. Longcore, C. Moritz, C. J. Briggs, and J. W. Taylor. 2007. Popula-
tion genetics of the frog-killing fungus Batrachochytrium dendrobatidis. Proceedings of the
National Academy of Sciences, USA 104(34):13845–13850.
Piotrowski, J. S., S. L. Annis, and J. E. Longcore. 2001. Physiology, zoospore behavior, and enzyme
production of Batrachochytrium dendrobatidis, a chytrid pathogenic to amphibians. Phytopa-
thology 91(6 Suppl):S121.
Pounds, J. A., M. P. L. Fogden, and J. H. Campbell. 1999. Biological response to climate change on
a tropical mountain. Nature 398(6728):611–615.
Ramsey, J. P., L. K. Reinert, L. K. Harper, D. C. Woodhams, and L. A. Rollins-Smith. 2010. Immune
defenses against Batrachochytrium dendrobatidis, a fungus linked to global amphibian declines,
in the South African clawed frog, Xenopus laevis. Infection and Immunity 78(9):3981–3992.
Retallick, R. W. R., H. McCallum, and R. Speare. 2004. Endemic infection of the amphibian chytrid
fungus in a frog community post-decline. PLoS Biology 2(11):1965–1971.
Rollins-Smith, L. A., J. E. Longcore, S. K. Taylor, J. C. Shamblin, J. M. Krepp, and C. Carey. 2000.
Antimicrobial peptides of frog skin: A possible defense against pathogens associated with global
amphibian declines. Developmental and Comparative Immunology 24(Suppl 1):S20.
Rosenblum, E. B., J. Voyles, T. J. Poorten, and J. E. Stajich. 2010. The deadly chytrid fungus: A story
of an emerging pathogen. PLoS Pathogens 6:e10.
Skerratt, L. F., L. Berger, R. Speare, S. Cashins, K. R. McDonald, A. D. Phillott, H. B. Hines, and N.
Kenyon. 2007. Spread of chytridiomycosis has caused the rapid global decline and extinction
of frogs. EcoHealth 4(2):125–134.
Stebbins, R. C., and N. W. Cohen. 1995. A natural history of amphibians. Princeton, NJ: Princeton
University Press.
Stuart, S., J. S. Chanson, N. A. Cox, B. E. Young, A. S. L. Rodrigues, D. L. Fishman, and R. W.
Waller. 2004. Status and trends of amphibian declines and extinctions worldwide. Science
306:1783–1786.
Voyles, J., S. Young, L. Berger, C. Campbell, W. F. Voyles, A. Dinudom, D. Cook, R. Webb, R. A.
Alford, L. F. Skerratt, and R. Speare. 2009. Pathogenesis of chytridiomycosis, a cause of cata-
strophic amphibian declines. Science 326:582–585.
Vredenburg, V. T. 2004. Reversing introduced species effects: Experimental removal of introduced
fish leads to rapid recovery of a declining frog. Proceedings of the National Academy of Sci-
ences, USA 101(20):7646–7650.
Vredenburg, V. T., G. Fellers, and C. Davidson. 2005. The mountain yellow-legged frog Rana
muscosa (camp 1917). In Status and conservation of U.S. amphibians, edited by M. Lanoo.
Berkeley: University of California Press. Pp. 563–566.
Vredenburg, V. T., R. Bingham, R. Knapp, J. A. T. Morgan, C. Moritz, and D. Wake. 2007. Concor-
dant molecular and phenotypic data delineate new taxonomy and conservation priorities for
the endangered mountain yellow-legged frog. Journal of Zoology (London) 271(4):361–374.
Vredenburg, V. T., R. A. Knapp, T. S. Tunstall, and C. J. Briggs. 2010a. Dynamics of an emerging
disease drive large-scale amphibian population extinctions. Proceedings of the National Acad-
emy of Sciences, USA 107(21):9689–9694.
Vredenburg, V. T., J. M. Romansic, L. M. Chan, and T. Tunstall. 2010b. Does UV-B radiation affect
embryos of three high elevation amphibian species in California? Copeia 2010(3):502–512.
Wake, D. B., and V. T. Vredenburg. 2008. Are we in the midst of the sixth mass extinction? A
view from the world of amphibians. Proceedings of the National Academy of Sciences, USA
105:11466–11473.
APPENDIX A 355
A19
Introduction
Accelerating losses of amphibian biodiversity were brought to the forefront
of conservation concerns in 1989 at the First World Congress of Herpetology,
and the National Research Council Workshop the following year (Blaustein and
Wake, 1995). Although obvious causes of amphibian declines were known (e.g.,
habitat destruction and alteration, introduction of exotic species, pollution), the
causes of some declines and disappearances were undetermined, such as enig-
matic declines in tropical Australia, Meso-America, and western United States.
Detailed retrospective studies of declines in California national parks (Drost and
Fellars, 1996) and Monteverde Cloud Forest Preserve in Costa Rica (Pounds
et al., 1997) gave conclusive evidence of community-wide declines. Amphibian
species have disappeared or declined at such an alarming rate over the past few
decades that the phenomenon is now considered to be an amphibian extinction
crisis (Mendelson et al., 2006; Wake and Vredenburg, 2008).
In 2004, the Global Amphibian Assessment (GAA) conducted by the Interna-
tional Union for Conservation of Nature (IUCN) revealed that 32.5 percent of the
world’s ca. 6,000 amphibian species are currently threatened with extinction, and
more than 120 have already disappeared (Stuart et al., 2004). Furthermore, half
of 435 rapidly declining species were threatened by enigmatic processes (e.g.,
disease, climate change) and disease was detected in a significant proportion of
families (15/36 families) with threatened species (Stuart et al., 2004). While long
suspected, little direct evidence for the emergence of infection existed until a new
pathogen to science, the fungal chytrid Batrachochytrium dendrobatidis (Bd),
was found in dead and dying frogs collected in the mid-1990s from declining
populations in Australia and Panama (Berger et al., 1998; Longcore et al., 1999).
59╛╛Unit for Environmental Research: Zoology, Private Bag X6001, North-West University, Potchef-
isms targeted for conservation (e.g., amphibians), the very group conservationists
are trying to protect need to be mitigated to prevent the pathogen from spreading.
The disease pathway is not always a single invasive host–pathogen system,
but may involve multiple host species that are either directly or indirectly related
to the source population. Several invasive frog species play a role in the exporta-
tion of Bd from South Africa and its ongoing spread and impact on amphibian
populations around the world. Trade in African clawed frogs (Xenopus laevis)
since the 1930s was identified as one of the first human-mediated pathways for
spreading Bd internationally (Weldon et al., 2004). Feral populations that be-
came established in countries where the species had been moved (Weldon et al.,
2007) allowed for potential transmission of Bd to naïve amphibian species. Such
a link was hypothesized to exist between X. laevis and the American bullfrog
Rana catesbeiana, an invasive species and major vector of the disease (Daszak
et al., 2004; Weldon et al., 2004). Various other strong links have been confirmed
between the international amphibian trade for use in the food, pet, and labora-
tory industry, and the global dispersal of Bd (Fisher and Garner, 2007; Garner
et al., 2006; Schloegel et al., 2009). A second species traded from South Africa,
Xenopus gilli, was identified as the source of a Bd infection transmitted to captive
Alytes muletensis in Spain (Walker et al., 2008). These infected toads were later
reintroduced into their native habitat on the island of Mallorca, thus introducing
the pathogen into this disease-naïve ecosystem. What makes this case even more
extraordinary is that Bd was transmitted from one critically endangered species
to another, dismissing the assumption that the more common, widely distributed
species are mostly responsible for disease spread. This case study demonstrates
the caution with which reintroduction programs should be approached when there
is a potential for disease organisms to be vectored alongside their hosts; it is now
clear that invasive frogs and toads have spontaneously exposed many native spe-
cies of amphibians around the world to Bd, often with catastrophic results.
is the Lacy Act of 1900 that identifies a priority list of invasive species requiring
conservation management. The federal agencies responsible for implementing
the law are the U.S. Fish and Wildlife Service (FWS) and the National Oceanic
and Atmospheric Association, but the FWS does not provide specific authority to
prohibit the importation of a pathogen.
International trade in species whose existence may be threatened as a result
of commerce and other trafficking is regulated by the Convention on International
Trade in Endangered Species of Wild Fauna and Flora (CITIES). However, com-
prehensive data covering all amphibians is hard to find because there is no global
database or monitoring system for the trade in non-CITIES species (Schlaepfer
et al., 2005). Consequently, control over the international spread of Bd is not
regulated under this process when the vectors are non-threatened species such as
X. laevis and R. catesbeiana. The World Organisation for Animal Health (OIE)
provides guidelines that can be used by member countries to protect themselves
from the introduction of diseases and pathogens. Only two amphibian pathogens
(Bd and Ranavirus) are currently listed as notifiable to the OIE, thus requiring
regulation of the amphibian trade aimed at the prevention of disease spread.
Because non-members are under no obligation to adhere to the guidelines set
by the OIE and because recommendations rely on voluntary participation and
lack any legally binding mechanisms, amphibians traded by both member and
non-member countries are often not accompanied by veterinary certification.
This shortfall in international legislation may cause ambiguity with respect to
biosecurity associated with trade, ultimately contributing to the unabated global
spread of pathogens.
A.
B.
FIGURE A19-1╇ Maps indicating (A) the global prevalence and (B) regional U.S. preva-
Figure asA19-1.eps
lence of Batrachochytrium dendrobatidis mapped by www.bd-maps.net.
SOURCE: The Global Bd-Mapping Project, www.bd-maps.net.
2 bitmaps w some vector type
Group, the World Association of Zoos and Aquariums, and the IUCN/SSC Am-
phibian Specialist Group initiated the Amphibian Ark (AArk), an organization
that provides all levels of support to ex situ actions around the world (Pavajeau
et al., 2008). Ex situ conservation centers, in addition to maintaining captive as-
surance populations, provide a great opportunity for much-needed research on
the management of wildlife diseases.
Some concerns that have to be addressed when initiating captive assurance
populations is the need of substantial resources for the construction of biosecure
facilities, training of keepers, and support of amphibian husbandry requirements
(Gagliardo et al., 2008; Pavajeau et al., 2008). In addition, for animals salvaged
from areas with recognized disease-mediated die-offs, it is necessary to control
the population-limiting infection to help ensure the success of the captive colony
and to prevent introduction of disease to other collection animals (Pessier, 2008).
ponds that had been desiccated during treatment, and (3) itraconazole baths in
situ in artificial pools combined with habitat improvement, which included desic-
cation before releasing the tadpoles. The different methods resulted in varying
success, although generally, released tadpoles became reinfected and metamorphs
had a higher survival rate than what was determined for previous seasons. It
seems that Bd eradication is difficult to achieve and that attempts at keeping
infection levels controlled in the hope of achieving natural selection against the
disease may be the only current approach for in situ mitigation (Bosch et al.,
2010). In tandem with antifungal mitigations, the use of probiotic amphibian-
associated bacteria holds promise for managing levels of chytrid infection in
nature. Specifically, Janthinobacterium lividum has shown to clear infections
in captive populations and is currently being widely trialed in a variety of field
settings (Harris et al., 2009).
According to Schlaepfer et al. (2005), management practices that rely on
continuous intervention are not sustainable indefinitely due to human and mon-
etary resource limitations. Instead, Schlaepfer et al. (2005) suggest a novel ap-
proach where native species can respond to changes in their selective regime via
evolution or learning that allows coexistence of native and introduced species in
cases where the eradication of invasive species is not successful. Although this
approach is suggested for invasive species that threaten local species through
predation or competition, its application for a pathogen like Bd—capable of
rapid population extirpations and even extinction through disease—remains to
be explored. Some evidence exists that amphibian populations that have expe-
rienced chytridiomycosis-associated decline may partially recover and persist
with endemic Bd infection (Retallick et al., 2004; Woodhams et al., 2007). These
observations raise hope that some species can acquire resistance to chytridiomy-
cosis in the wild, lessening the reliance on the development of survival-assurance
colonies for species threatened by this disease (Mendelson et al., 2006; Young
et al., 2007).
captivity, who are continually treated to eliminate infections, may suffer a disad-
vantage when released into the wild because of increased susceptibility to infec-
tion caused by relaxed selection for resistance, especially in traits that are costly
to maintain (Smith et al., 2009). Given this possibility, species management
programs, especially those that include captive breeding and reintroductions, may
need to focus on maintaining the levels of immunity or variation in resistance that
are present in natural populations (Altizer and Pedersen, 2008).
Summary
The diversity and apparent increase in wildlife diseases, including chytrid-
iomycosis, have raised concerns that pathogens may pose a substantial threat
to biodiversity. The amphibian extinction crisis has been held to represent the
greatest species conservation challenge in the history of humanity (Pavajeau
et al., 2008). In part, this crisis has been accelerated by the spread of Bd through
the global amphibian trade. The ability of some amphibians to become alien
invasive species while functioning as carriers of chytridiomycosis complicates
conservation management of the matter. Response to amphibian threats usually
involves introducing or rising protection measures of the species and/or their
habitats. However, dealing with chytridiomycosis in wildlife populations re-
quires novel strategies that conventional approaches to and policy for amphibian
conservation do not adequately provide. It requires greater collaboration among
wildlife ecologists, veterinarians, and conservation organizations to provide a
more comprehensive mitigation strategy that may significantly reduce the risk of
chytridiomycosis-mediated population declines.
The global conservation community responded to this need in the ACAP, of
which the AArk plays an integral part, to select species that would otherwise go
extinct for captive assurance populations until they can be secured in the wild
(Pavajeau et al., 2008). Such facilities are ideally positioned to combine the man-
agement of endangered species through assurance colonies with opportunities for
conservation research that may include the development of new approaches to
mitigate wildlife diseases. Future efforts to better manage and protect the planet’s
amphibian diversity will depend on innovation from both the sciences and the
world of policies and regulation. A wide spectrum of international and domestic
regulation regimes already play a significant role in shaping the conservation
status of amphibian biodiversity, but considerate input from the scientific com-
munity will make implementation more effective.
References
Altizer, S., and A. B. Pedersen. 2008. Host–pathogen evolution, biodiversity and disease risks for
natural populations. In Conservation biology: Evolution in action, edited by S. Carroll and C.
Fox. Oxford, UK: Oxford University Press. Pp. 259–278.
APPENDIX A 365
Anita, M., R. S. Thorpe, E. Hypolite, and A. James. 2007. A report on the status of the herpetofauna
of the Commonwealth of Dominica, West Indies. Applied Herpetology 4(2):177–194.
Berger, L., R. Speare, P. Daszak, D. E. Green, A. A. Cunningham, C. L. Goggin, R. Slocombe, M. A.
Ragan, A. D. Hyatt, K. R. McDonald, H. B. Hines, K. R. Lips, G. Marantelli, and H. Parkes.
1998. Chytridiomycosis causes amphibian mortality associated with population declines in the
rain forests of Australia and Central America. Proceedings of the National Academy of Sciences,
USA 95:9031–9036.
Berger, L., G. Marantelli, L. F. Skerratt, and R. Speare. 2005. Virulence of the amphibian chytrid
fungus Batrachochytrium dendrobatidis varies with the strain. Diseases of Aquatic Organisms
68:47–50.
Bielby, J., N. Cooper, A. A. Cunningham, T. W. J. Garner, and A. Purvis. 2008. Predicting susceptibil-
ity to future declines in the world’s frogs. Conservation Letters 1:82–90.
Blaustein, A. R., and D. B. Wake. 1995. The puzzle of declining amphibian populations. Scientific
American 272:52–57.
Bosch, J., S. Fernández-Beaskoetxea, and B. Martín-Beyer. 2010. Time for chytridiomycosis mitiga-
tion in Spain. Aliens: The Invasive Species Bulletin 30:54–58.
Briggs, C. J., R. A. Knapp, and V. T. Vredenburg. 2010. Enzootic and epizootic dynamics of the
chytrid fungal pathogen of amphibians. Proceedings of the National Academy of Sciences, USA
107:9695–9700.
Cunningham, A. A. 1996. Disease risks of wildlife translocations. Conservation Biology 10:349–353.
Cunningham, A. A., P. Daszak, and J. P. Rodriguez. 2003. Pathogen pollution: Defining a parasitologi-
cal threat to biodiversity conservation. Journal of Parasitology 89(Suppl):S78–S83.
Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2001. Anthropogenic environmental change and the
emergence of infectious disease in wildlife. Acta Tropica 78:103–116.
Daszak, P., A. Strieby, A. A. Cunningham, J. E. Longcore, C. C. Brown, and D. Porter. 2004. Experi-
mental evidence that the bullfrog (Rana catesbeiana) is a potential carrier of chytridiomycosis,
an emerging fungal disease of amphibians. Herpetological Journal 14:201–207.
Drost, C. A., and G. M. Fellars. 1996. Collapse of a regional frog fauna in the Yosemite area of the
California Sierra Nevada, USA. Conservation Biology 10(2):414–425.
Fenichel, E. P. and R. D. Horan. 2007. Jointly-determined ecological thresholds and economic trade-
offs in wildlife disease management. Natural Resource Modeling 20(4):511–547.
Fisher, M. C., and T. W. J. Garner. 2007. The relationship between the emergence of Batrachochytrium
dendrobatidis, the international trade in amphibians and introduced amphibian species. Fungal
Biology Reviews 21:2–9.
Fisher, M. C., T. W. J. Garner, and S. F. Walker. 2009a. The global emergence of Batrachochytrium
dendrobatidis and amphibian chytridiomycosis in space, time and host. Annual Review of Mi-
crobiology 63:291–310.
Fisher, M. C., J. Bosch, Z. Yin, D. A. Stead, J. Walker, L. Selway, A. J. Brown, L. A. Walker, N. A.
Gow, J. E. Stajich, and T. W. J. Garner. 2009b. Proteomic and phenotypic profiling of the am-
phibian pathogen Batrachochytrium dendrobatidis show that genotype is linked to virulence.
Molecular Ecology 18:415–426.
Gagliardo, R., P. Crump, E. Griffith, J. Mendelson, H. Ross, and K. Zippel. 2008. The principles of
rapid response for amphibian conservation, using the programmes in Panama as an example.
International Zoo Yearbook 42:125–135.
Garner, T. W. J., M. W. Perkins, P. Govindarajulu, D. Seglie, S. Walker, A. Cunningham, and M. C.
Fisher. 2006. The emerging amphibian pathogen, Batrachochytrium dendrobatidis, globally
infects introduced populations of the North American bullfrog, Rana catesbeiana. Biology
Letters 2:455–459.
Green, D. E., K. A. Converse, and A. Schrader. 2002. Epizootiology of sixty-four amphibian morbid-
ity and mortality events in the USA, 1996–2001. Annals of the New York Academy of Sciences
969:323–339.
366 FUNGAL DISEASES
Retallick, R. W. R., H. McCallum, and R. Speare. 2004. Endemic infection of the amphibian chytrid
fungus in a frog community post-decline. PLoS Biology 2:1965–1971.
Rödder, D., J. Kielgast, J. Bielby, S. Schmidtlein, J. Bosch, T. W. J. Garner, M. Veith, S. Walker, M. C.
Fisher, and S. Lötters. 2009. Global amphibian extinction risk assessment for the panzootic
chytrid fungus. Diversity 1:52–66.
Ron, S. R. 2005. Predicting the distribution of the amphibian pathogen Batrachochytrium dendroba-
tidis in the New World. Biotropica 37(2):209–222.
Schlaepfer, M. A., P. W. Sherman, B. Blossey, and M. C. Runge. 2005. Introduced species as evolu-
tionary traps. Ecology Letters 8:241–246.
Schloegel, L. M., J. M. Hero, L. Berger, R. Speare, K. McDonald, and P. Daszak. 2006. The decline
of the sharp-snouted day frog (Taudactylus acutirostris): The first documented case of extinction
by infection in a free-ranging wildlife species? EcoHealth 3:35–40.
Schloegel, L. M., A. Picco, A. M. Kilpatrick, A. Hyatt, and P. Daszak. 2009. Magnitude of the
U.S. trade in amphibians and presence of Batrachochytrium dendrobatidis and ranavirus in-
fection in imported North American bullfrogs (Rana catesbeiana). Biological Conservation
142:1420–1426.
Smith, K. F., K. Acevedo-Whitehouse, and A. B. Pedersen. 2009. The role of infectious diseases in
biological conservation. Animal Conservation 12:1–12.
Stuart, S., J. S. Chanson, N. A. Cox, B. E. Young, S. L. Rodrigues, D. L. Fischman, and R. W.
Waller. 2004. Status and trends of amphibian declines and extinctions worldwide. Science
306:1783–1786.
Stuart, S. N., M. Hoffmann, J. S. Chanson, N. A. Cox, R. J. Berridge, P. Ramani, and B. E. Young
(eds.). 2008. Threatened amphibians of the world. International Union for Conservation of
Nature. Gland, Switzerland, and Conservation International, Arlington, VA. Barcelona, Spain;
Lynx Editions.
Tucker, G. 2005. Biodiversity evaluation methods. In Handbook of biodiversity methods, edited by D.
Hill, M. Fasham, G. Tucker, M. Shewry, and P. Shaw. Cambridge, UK: Cambridge University
Press. Pp. 65–101.
Wake, D. B., and V. T. Vredenburg. 2008. Are we in the midst of the sixth mass extinction? A
view from the world of amphibians. Proceedings of the National Academy of Sciences, USA
105(Suppl):11466–11473.
Walker, S. F., J. Bosch, T. Y. James, A. P. Litvintseva, J. A. O. Valls, S. Piña, G. García, G. A. Rosa,
A. A. Cunningham, S. Hole, R. Griffiths, and M. C. Fisher. 2008. Invasive pathogens threaten
species recovery programs. Current Biology 18:R853–R854.
Wang, X., K. Zhang, Z. Wang, Y. Ding, W. Wu, and S. Huang. 2004 The decline of the Chinese giant
salamander Andrias davidianus and implications for its conservation. Oryx 38:197–202.
Webb, R., D. Mendez, L. Berger, and R. Speare. 2007. Additional disinfectants effective against the
amphibian chytrid fungus Batrachochytrium dendrobatidis. Diseases of Aquatic Organisms
74:13–16.
Weldon, C., L. H. Du Preez, A. D. Hyatt, R. Muller, and R. Speare. 2004. Origin of the amphibian
chytrid fungus. Emerging Infectious Diseases 10(12):2000–2005.
Weldon, C., A. L. De Villiers, and L. H. Du Preez. 2007. Quantification of the African clawed frog
trade from South Africa, with implications for biodiversity conservation. African Journal of
Herpetology 56:77–83.
Woodhams, D. C., R. A. Alford, and G. Marantelli. 2003. Emerging disease of amphibians cured by
elevated body temperature. Diseases of Aquatic Organisms 55:65–67.
Woodhams, D. C., V. T. Vrendenberg, M. Simon, D. Billheimer, B. Shakhtour, Y. Shyr, C. Briggs,
L. Rollins-Smith, and R. Harris. 2007. Symbiotic bacteria contribute to innate immune de-
fenses of the threatened mountain yellow-legged frog, Rana muscosa. Biological Conservation
138:390–398.
Young, S., L. Berger, and R. Speare. 2007. Amphibian chytridiomycosis: Strategies for captive man-
agement and conservation. International Zoo Yearbook 41:85–95.
368 FUNGAL DISEASES
A20
In North America, WNS is known to affect 6 species of bats that use hiber-
nation as their winter survival strategy: the big brown bat (Eptesicus fuscus),
the eastern small-footed bat (Myotis leibii), the little brown bat (M. lucifugus),
the northern long-eared bat (M. septentrionalis), the tricolored bat (Perimyotis
subflavus), and the Indiana bat (M. sodalis) (Blehert et al., 2009; Courtin et al.,
2010; Turner and Reeder, 2009). Since its detection in February 2006 in a popular
tourist cave near Albany, New York, USA, WNS has spread >1,300 km into Con-
necticut, Massachusetts, New Hampshire, New Jersey, Pennsylvania, Tennessee,
Vermont, Virginia, and West Virginia in the United States and the provinces of
Ontario and Quebec in Canada (Blehert et al., 2009, Turner and Reeder, 2009;
United States Geological Survey, 2010) in a pattern suggesting the spread of an
infectious agent.
A recently discovered psychrophilic (cold-loving) fungus, Geomyces de-
structans (Gargas et al., 2009), has consistently been isolated from bats that meet
the pathologic criteria for WNS, including colonization of skin by fungal hyphae
causing characteristic epidermal erosions and ulcers that can progress to invasion
of underlying connective tissue (Meteyer, et al.,2009; Reichard and Kunz, 2009).
G. destructans is identified by its distinctive asymmetrically curved conidia
and has a unique taxonomic position among other Geomyces spp. described to
date (Gargas et al., 2009) Its closest genetic relative is G. pannorum, a ubiqui-
tous psychrophilic, keratinolytic fungus that has been isolated from a variety of
sources and geographic regions, including soil and the fur of wild mammals in
France (Chabasse et al., 1987), floors of trains and ferryboats in Italy (Mercantini
et al., 1989), boreal forests in Canada (de Bellis et al., 2007), and environmental
samples from Arctic regions (Kochkina et al., 2007; Mercantini et al., 1989).
G. pannorum var. pannorum has been reported as an unusual dermatophyte
infecting fingernails and superficial skin of humans who have a history of close
contact with soil and dust (Gianni et al., 2003, Zelenková, 2006). However, G.
destructans differs from other common soil fungi of North America in its ability
to invade the living tissues of hibernating bats.
After WNS was described in North America (Blehert et al., 2009), reports
dating back to the early 1980s (Feldmann, 1984) described repeated observations
of white fungal growth on muzzles of hibernating bats in Germany. However,
these bats lacked the characteristics of WNS such as associated deaths. Moreover,
fungus was not identified. In response to WNS in North America, researchers in
Europe initiated a surveillance effort during the winter of 2008–09 for WNS-like
fungal infections among hibernating populations of bats in Europe. G. destruc-
tans in Europe was previously reported in 1 hibernating bat that was sampled in
France during March 2009 (Puechmaille et al., 2010).
In this report, we describe results of a more extensive effort by scientists
from 4 countries in Europe (Germany, United Kingdom, Hungary, and Switzer-
land) to obtain and analyze samples from hibernating bats with white patches
on their faces or wing membranes. Our objectives were to identify the fungus
370 FUNGAL DISEASES
colonizing such affected hibernating bats in Europe and to clarify its geographic
distribution over a broad area of Europe.
FIGURE A20-1╇ (A) Greater mouse-eared bat (Myotis myotis) with white fungal growth
around its muzzle, ears, and wing membranes (photograph provided by Tamás Görföl).
Figure A20-1AandB.eps
(B) Scanning electron micrograph of a bat hair colonized by Geomyces destructans. Scale
bar = 10 μm. bitmap
372 FUNGAL DISEASES
ing PCR with primers nu-SSU-0021–5′ (White et al., 1990) and nu-SSU-1750–3′
(Gargas and Taylor, 1992) as above, except the extension time was increased to
2 min. Sequencing primers were PCR primers; nu-SSU-0402–5′ (Gargas and
Taylor, 1992), nu-SSU-1150–5′ (White et al., 1990), nu-SSU-0497–3′ (Gargas
and Taylor, 1992), and nu-SSU-1184–3′ (Gargas et al., 1995) were added for
SSU. PCR products were sent to the Robert Koch Institute, Berlin, Germany, for
direct sequencing.
Culture analyses of samples were performed at Munich University Hospital
and IZW. After examining tape impressions by using light microscopy, we identi-
fied small areas with fungal conidia characteristic of G. destructans and excised
them with a sterile scalpel blade. Half of the excised material was used for PCR;
the remaining sample and samples of individual hairs with microscopic indication
of G. destructans were immediately placed onto Sabouraud dextrose agar plates
containing gentamicin and chloramphenicol and incubated at 4°C and 8°C. G.
destructans isolates obtained during this study are maintained at IZW.
Results
We obtained and analyzed samples from live bats with obvious fungal growth
on their bodies found between mid-February and the end of March 2009 at 11
sites (8 in Germany, 1 in Hungary, and 2 in Switzerland). Samples were also ob-
tained from an additional bat in Germany in February 2008 and from 2 dead bats
from a site in the United Kingdom in March 2009 (Tables A20-1, A20-2) All 12
hibernacula sampled contained 1–5 animals that exhibited obvious fungal growth.
Forty-three samples were obtained from these 12 hibernacula and represented 23
adult bats of 6 species: 1 Brandt bat (M. brandtii), 3 pond bats (M. dasycneme),
1 Daubenton bat (M. daubentonii), 1 lesser mouse-eared bat (M. oxygnathus), 15
greater mouse-eared bats (M. myotis), and 2 greater horseshoe bats (Rhinolophus
ferrumequinum).
After direct PCR amplification and DNA sequence analysis of fungal rRNA
gene ITS regions, genetic signatures 100% identical with those from G. destruc-
tans type isolate NWHC 20631–21 (GenBank accession no. EU884921) were
identified from 21 of 23 bats examined: 15/15 from Germany, 2/2 from Hungary,
and 4/4 from Switzerland. Both bats from the United Kingdom were colonized
by Penicillium sp. (Tables A20-1, A20-2). Fungi with conidia morphologically
identical to those of G. destructans (Figure A20-1, panel B) as described by
Gargas et al. (2009) were isolated in axenic cultures from 8 of 23 bats examined:
3/15 from Germany, 1/2 from Hungary, and 4/4 from Switzerland) (Tables A20-1,
A20-2; Figure A20-2).
Consistent with published descriptions for G. destructans (Gargas et al.,
2009), fungal colonies grew slowly and within 14 days attained diameters of 1.0
mm at 4°C and 4.0–5.0 mm at 8°C; no growth occurred at 25°C. The sensitiv-
ity of our method for isolating G. destructans from bat hair was comparable to
APPENDIX A 373
published diagnostic sensitivity for culturing G. destructans from bat skin (Lorch
et al., 2010). Subsequent PCR/DNA sequencing analyses of the 8 isolates indi-
cated that they all had rRNA gene ITS and SSU region DNA sequences identical
to those of G. destructans type isolate NWHC 20631–21 (GenBank accession
nos. EU884921 for ITS and FJ231098 for SSU).
Unlike other bats sampled in this study, the 2 greater horseshoe bats from the
United Kingdom were found dead, and their nostrils were colonized by Penicil-
lium sp. These bats did not fulfill the pathologic criteria for WNS (Meteyer, et al.,
2009) because fungal hyphae did not invade the epidermis but remained within
the superficial layer of the epidermal stratum corneum. A more complete descrip-
tion of the postmortem analysis of the greater horseshoe bats has been reported
(Barlow et al., 2009). G. destructans was not isolated in culture, and its genetic
signature was not identified by PCR and DNA sequencing of samples collected
from greater horseshoe bats.
Discussion
Laboratory analyses demonstrated that 5 species of the genus Myotis in
Europe harbored G. destructans; male and female bats were equally affected.
Despite laboratory confirmation that bats obtained in Germany, Switzerland, and
Hungary were colonized by G. destructans, deaths were not observed at collec-
tion sites. Puechmaille et al. (2010) reported a similar observation with a greater
mouse-eared bat in France. Additionally, a lesser mouse-eared bat from Hungary
with visible fungal infection during hibernation, from which G. destructans was
isolated, was recaptured 5 months later (August 2009) and showed no external
signs of fungal infection. On February 19, 2010, the same bat was again observed
in the same hibernaculum without any visible sign of fungal growth. However, 7
other bats within that group of 55 animals displayed obvious fungal growth but
were not sampled for this study.
TABLE A20-2╇ Fungal Culture and PCR Results for 23 Bats with Evidence of Fungal Colonization Tested by Light or
374
FIGURE A20-2╇ Locations in Europe of bats positive for Geomyces destructans by PCR
alone (circles) or by PCR and culture (solid stars) and bats negative for G. destructans
Figure A20-2.eps
but positive for other fungi (square). Numbers for locations correspond to those in Table
bitmap
A20-2. Sites 7, 8, and 9 had additional bats that were positive for G. destructans only by
PCR. Location of a bat positive for G. destructans in France (Puechmaille et al., 2010)
is indicated by an open star. Some sites had >1 bat species with evidence of colonization
by G. destructans.
indicating that the fungus is most likely spread as local bat populations emerge
from hibernacula, disperse, and interact with populations within their dispersal
range. Identification of bats colonized by G. destructans from such distant sites,
in addition to the relatively homogenous distribution of the fungus among sites
in Germany, suggests that G. destructans may be widespread in Europe.
Regardless of widespread occurrence of G. destructans among bat species
in Europe (Figure A20-2), deaths of bats in Europe caused by WNS, similar to
those caused by WNS in North America, have not been observed. Although no
bat species migrates between Europe and North America or is present on both
continents (Dietz et al., 2009; Nowak, 1999), many species of the genus Myotis
are infected by G. destructans on each continent. Although the mechanism(s) by
which hibernating bats died because of infection with G. destructans in North
America is not yet understood, bat species in Europe may exhibit greater resis-
tance or respond differently to infection by this fungus than their counterparts in
North America.
Before the emergence of WNS in North America, large aggregations of
hibernating bats ranging from 1,000 to 50,000 animals were common in caves
and mines of affected regions, and many hibernation sites in regions of North
America still unaffected by WNS contained tens of thousands of bats during
winter (some contain hundreds of thousands) (Barbour and Davis, 1969). In
contrast, aggregations of bats hibernating in caves and mines in Europe rarely
exceed 1,000 animals. However, larger hibernating groups have been observed
at a few natural sites, such as a cave in northern Germany with 13,000–18,000
bats (Petermann and Boye, 2006) and human-made structures (e.g., Daubenton
bats in bunkers and catacombs) (Dietz et al., 2009). If host density plays a role
in G. destructans transmission or deaths of bats, such as through increased dis-
turbance of clustered bats, the bats in Europe may experience lower mortality
rates because they form smaller hibernation groups composed of small clusters
or individual bats. Apparent continental differences in susceptibility of hibernat-
ing bats to deaths associated with skin infection by G. destructans may indicate
either circumstantial or evolved resistance in bats in Europe.
G. destructans has been detected in North America only in states and prov-
inces where WNS has also been observed and in contiguous states. Recent emer-
gence and spread of G. destructans with associated deaths of bats throughout
hibernacula in the northeastern United States (Turner and Reeder, 2009) may
suggest ecologic release of an exotic pathogen into an uninfected ecosystem.
Although this suggestion remains a hypothesis and how G. destructans may have
been introduced to the United States is not known, initial documentation of WNS
in a popular tourist cave near Albany, New York (Blehert et al., 2009), suggests
that a human vector could have been involved.
There are many examples of unintended introductions of fungal pathogens,
particularly of those affecting plants and ectothermic animals with tissue temper-
atures permissive to fungal infection (Casadevall, 2005; Desprez-Loustau et al.,
APPENDIX A 377
2007; Robert and Casadevall, 2009). One case with striking similarities is the
panzootic chytrid fungus (Batrachochytrium dendrobatidis), which has caused
global decreases among amphibian species (Fisher et al., 2009). As with skin in-
fection by B. dendrobatidis in amphibians, which can alter body electrolyte levels
and lead to cardiac arrest (Voyles et al., 2009), skin infection by G. destructans
in hibernating bats may also kill by causing irreversible homeostatic imbalance
because wing membranes play major roles in water balance, circulation, and
thermoregulation of hibernating bats during winter (Davis, 1970; Makanya and
Mortola, 2007).
Bat species in Europe may be immunologically or behaviorally resistant
to G. destructans because of having coevolved with the fungus. Additionally,
microbial flora of bat skin or other abiotic surfaces in bat hibernacula in Europe
may have also coevolved to incorporate G. destructans as a nonpathogenic com-
ponent of the microbial community. Conversely, possible recent introduction of
G. destructans into the United States, with subsequent infection of bat species in
North America and ecosystems not infected with the fungus, provides a potential
explanation for the devastating effects of WNS in North America. Although bats
are reservoirs of various pathogens (Calisher et al., 2006; Wibbelt et al., 2009),
research into the immune function of bats, particularly during hibernation, is just
beginning.
In conclusion, nondetrimental colonization of bat species in Europe by G.
destructans may be relatively common (Figure A20-2), and historical reports
(Feldmann, 1984) suggest that such colonization of hibernating bats in Europe
has occurred for several decades. In contrast to recent mass deaths associated
with G. destructans skin infection, which is the hallmark of WNS in North
America, bats in Europe appear to coexist with G. destructans. Studies to inves-
tigate mechanisms of pathogenesis, microbial ecology, and phylogeography of G.
destructans will be essential for developing a comprehensive understanding of
WNS. In particular, testing the hypotheses that bats in Europe are more resistant
to fungal skin infection by G. destructans, that G. destructans was introduced
from Europe to North America, and that environmental circumstances limit the
pathogenicity of G. destructans in Europe seem warranted. Divergent manifesta-
tions of infection by G. destructans in bats in Europe and North America provide
a unique opportunity to address these research objectives with the ultimate goals
of better understanding WNS and developing sound strategies to manage the
devastating effects of this emerging wildlife disease in North America.
Acknowledgments
We thank N. Jahn, D. Viertel, A. Kus, and C. Kohl for excellent technical
assistance; A. Beck, T. Filip, M. Franz, S. Gloor, R. Güttinger, A. Kiefer, V. Korn,
G. Mäscher, B. Máté, C.D. Morawitz, C. Morris, P. Paulovics, M. Piskorski, F.
Putzmann, W. Schorcht, and C. Tress for help surveying sites and retrieving sam-
378 FUNGAL DISEASES
References
Barbour RW, Davis WH. Bats of America. Lexington (KY): The University Press of Kentucky; 1969.
Barlow A, Ford S, Green R, Morris C, Reaney S. Investigations into suspected white-nose syndrome
in two bats in Somerset. Vet Rec. 2009;165:481–2.
Blehert DS, Hicks AC, Behr M, Meteyer CU, Berlowski-Zier BM, Buckles EL, et al. Bat white-nose
syndrome: an emerging fungal pathogen? Science. 2009;323:227. DOI: 10.1126/science.1163874
Calisher CH, Childs JE, Field HE, Holmes KV, Schountz T. Bats: an important reservoir host of
emerging viruses. Clin Microbiol Rev. 2006;19:531–45. PubMed DOI: 10.1128/CMR.00017-06
Casadevall A. Fungal virulence, vertebrate endothermy, and dinosaur extinction: is there a connec-
tion? Fungal Genet Biol. 2005;42:98–106. DOI: 10.1016/j.fgb.2004.11.008
Chabasse D, Guiguen C. Couatarmanac’h A, Launay H, Reecht V, de Bièvre C. Keratinophilic fungal
flora isolated from small wild mammals and rabbit-warren in France. Discussion on the fungal
species found [in French]. Ann Parasitol Hum Comp. 1987;62:357–68.
Courtin F, Stone W, Risatti G, Gilbert K, Van Kruiningen H. Pathologic findings and liver el-
ements in hibernating bats with white-nose syndrome. Vet Pathol. 2010;47:214–9. DOI:
10.1177/0300985809358614
Davis WH. Hibernation: ecology and physiological ecology. In: Wimsatt WA, editor. Biology of bats.
Vol. 1. New York: Academic Press; 1970. p. 265–300.
de Bellis T, Kernaghan G, Widden P. Plant community influences on soil microfungal assemblages
in boreal mixed-wood forests. Mycologia. 2007;99:356–67. DOI: 10.3852/mycologia.99.3.356
Desprez-Loustau ML, Robin C, Buée M, Courtecuisse R, Garbaye J, Suffert F, et al. The fun-
gal dimension of biological invasions. Trends Ecol Evol. 2007;22:472–80. DOI: 10.1016/j.
tree.2007.04.005
Dietz C, von Helversen O, Nill D. Bats of Britain, Europe and North-west Africa. London: A and C
Black Publishers; 2009.
Feldmann R. Teichfledermaus–Myotis dasycneme (Boie, 1825). Die Säugetiere Westfalens. Münster:
Westfälisches Museum für Naturkunde; 1984. pp. 107–11.
Fisher MC, Garner TW, Walker SF. Global emergence of Batra-chochytrium dendrobatidis and
amphibian chytridiomycosis in space, time, and host. Annu Rev Microbiol. 2009;63:291–310.
DOI: 10.1146/annurev.micro.091208.073435
Gargas A, dePriest P, Taylor J. Positions of multiple insertions in SSU rDNA of lichen forming fungi.
Mol Biol Evol. 1995;12:208–18.
Gargas A, Taylor J. Polymerase chain reaction (PCR) primers for amplifying and sequencing 18S
rDNA from lichenized fungi. Mycologia. 1992;84:589–92. DOI: 10.2307/3760327
Gargas A, Trest MT, Christensen M, Volk TJ, Blehert DS. Geomyces destructans sp. nov. associated
with bat white-nose syndrome. Mycotaxon. 2009;108:147–54.
Gianni C, Caretta G, Romano C. Skin infection due to Geomyces pannorum var. pannorum. Mycoses.
2003;46:430–2. DOI: 10.1046/ j.1439-0507.2003.00897.x
Hutterer R, Ivanova T, Meyer-Cords C, Rodrigues L. Bat migrations in Europe: a review of banding
data and literature. Bonn (Germany): German Agency for Nature Conservation; 2005.
APPENDIX A 379
Kochkina GA, Ivanushkina NE, Akimov VN, Gilichinskii DA, Ozerskaia SM. Halo- and psychrotol-
erant Geomyces fungi from arctic cryopegs and marine deposits [in Russian]. Mikrobiologiia.
2007;76:39–44.
Lorch JM, Gargas A, Meteyer CU, Berlowski-Zier BM, Green DE, Shearn-Bochsler V, et al. Rapid
polymerase chain reaction diagnosis of white-nose syndrome in bats. J Vet Diagn Invest.
2010;22: 224–30.
Makanya AN, Mortola JP. The structural design of the bat wing web and its possible role in gas ex-
change. J Anat. 2007;211:687–697. PubMed DOI: 10.1111/j.1469-7580.2007.00817.x
Mercantini R, Marsella R, Cervellati M. Keratinophilic fungi isolated from Antarctic soil. Mycopatho-
logia. 1989;106:47–52. DOI: 10.1007/BF00436926
Mercantini R, Marsella R, Prignano G, Moretto D, Marmo W, Leonetto F, et al. Isolation of keratino-
philic fungi from the dust of ferry boats and trains in Italy. Mycoses. 1989;32:590–4.
Meteyer CU, Buckles EL, Blehert DS, Hicks AC, Green DE, Shearn-Bochsler V, et al. Histopatho-
logic criteria to confirm white-nose syndrome in bats. J Vet Diagn Invest. 2009;21:411–4.
Nowak R. Walker’s mammals of the world. Baltimore: The Johns Hopkins University Press; 1999.
Petermann R, Boye P. National report on bat conservation in the Federal Republic of Germany
2003–2006. Bonn (Germany): Eurobats; 2006.
Puechmaille SJ, Verdeyroux P, Fuller H, Ar Gouilh M, Bekaert M, Teeling EC. White-nose syndrome
fungus (Geomyces destructans) in bat, France. Emerg Infect Dis. 2010;16:290–3.
Reichard JD, Kunz TH. White-nose syndrome inflicts lasting injuries to the wings of little brown myotis
(Myotis lucifugus). Acta Chiropterologica. 2009;11:457–64. DOI: 10.3161/150811009X485684
Robert VA, Casadevall A. Vertebrate endothermy restricts most fungi as potential pathogens. J Infect
Dis. 2009;200:1623–6. DOI: 10.1086/644642
Turner GR, Reeder DM. Update of white-nose syndrome in bats, September 2009. Bat Research
News. 2009;50:47–53.
United States Geological Survey. Update on white-nose syndrome: Tennessee finding. USGS wildlife
health bulletin. Reston (VA): The Survey; 2010 [cited 2010 May 19]. http://www.nwhc.usgs.
gov/ disease_information/white-nose_syndrome/
Voyles J, Young S, Berger L, Campbell C, Voyles WF, Dinudom A, et al. Pathogenesis of chytridio-
mycosis, a cause of catastrophic amphibian declines. Science. 2009;326:582–5. DOI: 10.1126/
science.1176765
White T, Bruns T, Lee S, Taylor J. Amplification and direct sequencing of fungal ribosomal RNA
genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TH, editors. PCR proto-
cols: a guide to methods and applications. San Diego (CA): Academic Press; 1990. p. 315–22.
Wibbelt G, Speck S, Field H. Methods for assessing diseases in bats. In: Kunz T, Parsons S, editors.
Ecological and behavioral methods for the study of bats. Baltimore: Johns Hopkins University
Press; 2009. p. 775–94.
Zelenková H. Geomyces pannorum as a possible causative agent of dermatomycosis and onychomy-
cosis in two patients. Acta Dermatovenerol Croat. 2006;14:21–5
380 FUNGAL DISEASES
A21
63╛╛School of Biology and Environmental Science, University College Dublin, Dublin, Ireland
64╛╛Conway Institute of Biomolecular and Biomedical Research, Dublin, Ireland
65╛╛Email: s.puechmaille@gmail.com
66╛╛These authors contributed equally to this work.
67╛╛Leibniz Institute for Zoo and Wildlife Research, Berlin, Germany
68╛╛Office for Faunistic and Landscape Ecology, Schöneberg, Germany
69╛╛Plecotus Working Group,Association Natagora, Brussels, Belgium
70╛╛Robert Koch Institute, Berlin, Germany
71╛╛Museum and Institute of Zoology, Polish Academy of Sciences, Warszawa, Poland
72╛╛Commission de Protection des Eaux, du Patrimoine, de l’Environnement, du Sous-sol et des
France
80╛╛Department of Biology, Center for Bat Conservation in Northern Bavaria, Erlangen University,
Erlangen, Germany
81╛╛Nature and Biodiversity Conservation Union Rhineland-Palatine, Birkenfeld, Germany
82╛╛Bretagne Vivante SEPNB, Roussimel, Glénac, France
83╛╛Estonian Fund for Nature, Tartu, Estonia
84╛╛Sicista Development Centre, Tartu, Estonia
85╛╛Bat Working Group, Natuurpunt VZW, Belgium
86╛╛Nature and Biodiversity Conservation Union Southern Lower-Saxony,Nordstemmen, Germany
87╛╛Biotope Mapping Cooperation, Herford, Germany
88╛╛Institute of Natural Fibres and Medicinal Plants, Poznan, Poland
89╛╛Saxonian State Office for Environment Agriculture and Geology, Dresden-Pillnitz, Germany
â•… *Originally printed as: Sébastien Peuchmaille, Gudrun Wibbelt, Vaness Korn, Hubert Fuller,
Frédéric Forget, et al. �����������������������������������尓�����������������������������������
(2011) Pan-European Distribution of White-Nose Syndrome Fungus (Geomy-
ces destructans) Not Associated with Mass Mortality. PLoS ONE. 6(4): e19167. doi:10.1371/journal.
pone.0019167.
APPENDIX A 381
Abstract
Background:
The dramatic mass mortalities amongst hibernating bats in Northeastern
America caused by ‘‘white nose syndrome’’(WNS) continue to threaten popula-
tions of different bat species. The cold-loving fungus, Geomyces destructans,
is the most likely causative agent leading to extensive destruction of the skin,
particularly the wing membranes. Recent investigations in Europe confirmed
the presence of the fungus G. destructans without associated mass mortality in
hibernating bats in six countries but its distribution remains poorly known.
Methodology/Principal Findings:
We collected data on the presence of bats with white fungal growth in 12
countries in Europe between 2003 and 2010 and conducted morphological and
genetic analysis to confirm the identity of the fungus as Geomyces destructans.
Our results demonstrate the presence of the fungus in eight countries spanning
over 2000 km from West to East and provide compelling photographic evidence
for its presence in another four countries including Romania, and Turkey. Further-
more, matching prevalence data of a hibernaculum monitored over two consecu-
tive years with data from across Europe show that the temporal occurrence of
the fungus, which first becomes visible around February, peaks in March but can
still be seen in some torpid bats in May or June, is strikingly similar throughout
Europe. Finally, we isolated and cultured G. destructans from a cave wall adja-
cent to a bat with fungal growth.
Conclusions/Significance:
G. destructans is widely found over large areas of the European continent
without associated mass mortalities in bats, suggesting that the fungus is native to
Europe. The characterisation of the temporal variation in G. destructans growth
on bats provides reference data for studying the spatio-temporal dynamic of the
fungus. Finally, the presence of G. destructans spores on cave walls suggests that
hibernacula could act as passive vectors and/or reservoirs for G. destructans and
therefore, might play an important role in the transmission process.
Introduction
White nose-syndrome (WNS) is a devastating disease causing mass mortali-
ties in hibernating bats in North-America. In May 2009, it was estimated that
over one million bats had died from the disease which was first documented
in February 2006 at Howe’s Cave, West of Albany, New York (Anonymous,
2009). A visually conspicuous white fungus grows on the face, ears, or wings of
382 FUNGAL DISEASES
stricken bats with hyphae penetrating deep into the connective tissue of glabrous
skin and snout (Meteyer et al., 2009) and causing severe damage (Reichard and
Kunz, 2009). The fungus associated with WNS is a newly described, psychro-
philic (cold-loving) species (Geomyces destructans) (Gargas et al., 2009), closely
related to other psychrophilic species of Geomyces (Rice and Currah, 2006;
Puechmaille et al., 2010). Although it is not yet conclusively proven whether G.
destructans is the causative agent of the disease or if other co-factors are neces-
sary for disease to occur, the fungus is always found on bats at WNS sites where
hibernating bats experience mass mortalities (Blehert et al., 2009). To date, bac-
teriological, virological, parasitological and pathological evaluations as well as
studies of toxic contaminants have not identified the consistent presence of any
other agents/cause of death. The lack of evidence for the involvement of other
agents or compounds reinforces the suspicion that G. destructans is the causative
agent of WNS mortality (Meteyer et al., 2009; Blehert et al., 2009; Kannan et al.,
2010; Courtin et al., 2010).
Geomyces destructans has been found in nine species of bats in North-
America, from the provinces of Ontario and Quebec in Canada south and west
to the states of Tennessee and Oklahoma in the USA (Anonymous, 2010). Three
recent studies investigating samples collected in 2008–2010 have shown that
G. destructans was also present in six European countries (France, Germany,
Switzerland, Czech Republic, Slovakia & Hungary) (Puechmaille et al., 2010;
Martínková et al., 2010; Wibbelt et al., 2010). Nevertheless, the geographic
coverage of these studies was limited and the extent of the distribution of G. de-
structans in Europe remains poorly known. In this paper, we combine previously
published data on the distribution of G. destructans in Europe (Puechmaille et al.,
2010; Martínková et al., 2010; Wibbelt et al., 2010) with new data from twelve
countries covering 2,400 km from West to East (France to Turkey) and 1,900 km
from North to South (Estonia to Turkey) to demonstrate the widespread presence
of G. destructans on multiple species of hibernating bats in Europe without as-
sociated mass mortality.
Results
Switzerland and Slovakia, the fungus has been confirmed from 1–2 location(s) per
country, whereas it has been confirmed at 8 sites in Germany and 23 sites in the
Czech Republic (Puechmaille et al., 2010; Martínková et al., 2010; Wibbelt et al.,
2010). All confirmed detections of G. destructans in Europe have been made by
isolating and/or genetically identifying the fungus from hairs, swabs or touch im-
prints from bats (Puechmaille et al., 2010; Martínková et al., 2010; Wibbelt et al.,
2010). In Europe, eight species of Myotis have been observed being colonised by
G. destructans: M. Myotis, M. blythii (referred to as M. oxygnathus in [Wibbelt
et al., 2010]), M. mystacinus, M. daubentonii, M. dasycneme, M. nattereri, M.
bechsteinii and M. brandtii. Species from other bat families were present in the
caves with infected individuals (e.g., Miniopteridae: Miniopterus schreibersii;
Rhinolophidae: Rhinolophus hipposideros and R. ferrumequinum),but no G. de-
structans has been confirmed from these species. Previous extensive surveys
of cave fungi in Europe (i.e., Nováková, 2009; Mosca and Campanino, 1962;
Bastian et al., 2009) or fungi associated with insects hibernating in underground
sites (Kubátová and Dvořák, 2005) never reported G. destructans in their inven-
tory, although some other species of Geomyces were recovered (Nováková, 2009;
Mosca and Campanino, 1962; Bastian et al., 2009).
pone.0019167.t001
viously isolated (see list above). In some cases, we were not able to discriminate
between M. Myotis and M. blythii (referred to as M. Myotis/blythii) as well as
between the newly recognised M. escalerai (Ibañez et al., 2006; Cabrera, 1904)
and Myotis sp. A (Garcia-Mudarra et al., 2009), a yet undescribed cryptic species
from the M. nattereri species complex (Ibañez et al., 2006; Mayer et al., 2007).
Additionally, swab samples were collected from the tunnel wall of an Estonian
hibernaculum. On the 23rd May 2010, a M. brandtii was observed in this hiber-
naculum with white fungal growth on its snout (Figure A21-2A) but no sample
was collected at the time. When the site was revisited for sample collection on the
1st of June 2010, the bat had left the site so samples were collected by swabbing
the wall of the tunnel where the bat was seen nine days before. Four cotton swabs
were used to sample different areas a few centimetres around the location where
APPENDIX A 385
the bat was observed. The four swabs were then streaked onto four Sabouraud’s
agar plates each and monitored regularly to physically remove any fungal growth
that was not similar to G. destructans. Although the amount of fungi varied per
swab sample, G. destructans was recovered from all four swabs, henceforth con-
sidered as a single sample, bringing the total of samples analysed to 23. No mass
mortality was reported at any of the sites investigated.
386 FUNGAL DISEASES
October 13th) (Puechmaille et al., 2010; Blehert et al., 2009, Wibbelt et al., 2010;
Chaturvedi et al., 2010).
Figure A21-2.eps
bitmap
APPENDIX A 389
FIGURE A21-2╇ (facing page). Photographic evidence showing bats with confirmed or
suspected growth of G. destructans. Photographs of cases confirmed by genetic analysis,
from (A) Estonia (M. brandtii, May 23rd 2010, © L. Lutsar), (B) Poland (M. myotis,
March 7th 2010, © A. Wojtaszewski), (C) Germany (M. myotis, March 10th 2010, © C.
Jungmann), (D) France (M. myotis, March 4th 2010, © Y. Le Bris), (E) Netherlands (M.
daubentonii, March 9th 2010, © T. Bosch), (F) Germany (M. myotis, March 23rd 2010,
© K. Passior) (G) Belgium (M. mystacinus, March 18th 2010, © B. Mulkens), (H) Ger-
many (M. mystacinus, March 23rd 2010, © K. Passior) or bats with white-fungal growth
suspected as G. destructans from (I) Denmark (M. dasycneme, March 14th 2010, © B.
Ohlendorf), (J) Austria (M. myotis, February 2nd 2007, © O. Gebhardt), (K) Hungary (M.
myotis, February 19th 2010, © T. Görföl), (L) Belgium (M. myotis, March 7th 2010, © F.
Forget), (M) France (M. myotis, February 13th 2010, © J. Vittier), (N) Ukraine (M. myotis,
February 13th 2010, © A.-T. Bashta), (O) France (M. escalerai/sp. A, June 25th 2010, © F.
Blanc), (P) Turkey (M. myotis/blythii, March 22nd 2009, © M. Doker), and (Q) Romania
(M. blythii, March 29th 2008, © B. Szilárd).
doi:10.1371/journal.pone.0019167.g002
FIGURE A21-3╇ Seasonal changes of the number of live bats reported with white fungal
growth in Europe. The number of bats with visible white fungal growth at an hibernacu-
lum in Germany was monitoredFigureduring theA21-3.eps
winter 2006/2007 (blue line) and the winter
2007/ 2008 (green line). The vertical redbitmap
lines represent the number of Gd-suspect bats (or
confirmed) observed across twelve European countries (n = 127) from 2003 until 2010.
In the X-axis, the thick marks represent the start of each month. doi:10.1371/journal.
pone.0019167.g003
Discussion
Kubátová & Dvořák (2005) investigated fungi associated with insects hibernating
in underground sites but did not find Geomyces species. To our knowledge, only
one study in Europe has investigated fungi present in bats’ skin and hair samples
where, based on our current knowledge, G. destructans is most likely to be found.
During the winter 1999/2001, Larcher et al. (2003) collected 25 samples of hair
and skin swabs from six species, including three Myotis Myotis, but did not find
any Geomyces species. It is important to note that most fungal cultures have been
carried out at temperatures above 24–25°C, temperatures at which G. destructans
does not grow (Gargas et al., 2009; Chaturvedi et al., 2010), which could explain
why although present, this fungal species had never been reported in Europe
before the study of Puechmaille et al. (2010).
Combining previously published data from France, Germany, Switzerland,
Hungary, The Czech Republic and Slovakia (Puechmaille et al., 2010; Martínková
et al., 2010; Wibbelt et al., 2010), additional data collected from France, Germany
and Hungary (this study), and new data from Belgium, The Nether-lands, Poland,
Estonia and Ukraine (this study), we demonstrate here that G. destructans is
widespread in Europe. We consider the photographic evidence of bats with white
fungus matching the characteristic growth pattern (e.g., Figure A21-2; pictures
from Romania and Turkey) to most likely represent G. destructans, because so far
all tested live European bats with such white fungal growth on their nose, simi-
lar to Figure A21-2, have been confirmed to carry that species of fungus. These
findings further support the fact that G. destructans is widespread across Europe.
However, to confirm the presence of G. destructans in Europe prior to 2008, his-
torical collections of bat specimens (or cave soil samples), especially specimens
collected during the hibernation period, should be screened for the fungus.
As depicted in Figure A21-1, most cases of bats with G. destructans (con-
firmed and suspected) have been found from North-eastern France through Bel-
gium, The Netherlands, Germany and the Czech Republic. However, it is not
clear whether this pattern reflects an actual higher occurrence and/or prevalence
of the fungus in these regions or if it is at least partly due to sampling bias,
whereby the fungus is more likely to be detected in regions with a higher number
of underground sites visited every winter or in regions were the fungus is specifi-
cally sought. In our opinion, it is most likely that this large-scale pattern is due to
a sampling bias. For example, the largest number of sites with G. destructans in
any European country was reported from the Czech Republic (76 localities with
suspected or confirmed G. destructans) were most sites have been searched for
signs of the fungus (>800 hibernacula) (Martínková et al., 2010).
during the hibernation period or that the fungus is carried by bats at the onset of
hibernation but needs time to develop the visible white fungal growth due to the
phenology of the fungus. Therefore, the absence of visible white fungal growth
on bats when observed with the naked eye may not directly reflect the absence
of G. destructans, but rather just the absence of visible fungal colonies. Further
complicating matters, our ability to detect G. destructans growth on bats can
substantially differ with proximity to the bats (i.e., low ceiling versus high ceil-
ing) or the location of the bat (ceiling versus crevices).
Our results confirm the suggestion of Martínková et al. (2010) by showing
that during the hibernation period, bats can remove the fungus from their snout,
ears and wings to a point where the fungus is no longer visible to the naked eye,
although some spores might still be present on their skin. During hibernation,
bats arouse every two weeks on average (Brack and Twente, 1985; Twente et al.,
1985) and if bats consistently groom off the fungus on these occasions, our ability
to visually detect the fungus, if present, will be considerably reduced. We also
showed that towards the end of the hibernation period, bats were emerging from
the hibernaculum without visible signs of the fungus despite showing visible
white fungal growth from two weeks to five days before leaving the hibernacu-
lum. It would be important to investigate whether bats carry spores out of hiber-
nacula and as a result could possibly contaminate maternity roosts and maternity
mates as suggested by Hallam and McCracken (2011).
conditions have a direct or indirect effect on the growth rates, prevalence, and
detectability of G. destructans on bats.
It is crucial that the change in prevalence or detectability over the hiberna-
tion period is considered when comparing prevalence across sites and/or years.
Our results from monitoring one site throughout the hibernation period over
two consecutive years as well as reported cases from multiple sites in multiple
years show that bats with fungal growth are first seen in January, the number of
cases slowly increases into February and peaks in March, then in April when
bats emerge from hibernation it drops again. Our results are in agreement with
recent results from the Czech Republic where in the winter 2009/2010, the num-
ber of sites with bats with white fungal growth increased from 4.1% in January/
February (33/800 sites; regular bat monitoring) to 77.5% in late February/March
(76/98 sites; additional inspections) (Martínková et al., 2010) The Czech study
reported that this increase in G. destructans prevalence was ‘‘suggestive of an epi-
zootic spread of the fungus’’ (Martínková et al., 2010); we propose an alternative
explanation whereby the increase in prevalence of G. destructans in late winter
(March) might regularly (yearly) occur in Europe but has gone unnoticed. Nearly
all hibernation counts in previous years were carried out between December and
mid-February when prevalence/detectability of G. destructans is low, but not in
March (Battersby, 2010) when the prevalence/detectability of G. destructans is at
its highest (Figure A21-3). Although the total numbers of bats in the hibernacula
decreased through April as bats left for the maternity colonies, our results show
that there is a high probability of fungal growth developing on the remaining
individuals. This further supports our hypothesis proposed above and links the
duration of the hibernation period with the prevalence of G. destructans. By
increasing the sample size, some cases might be reported earlier in the hiberna-
tion season or later through the summer, but we expect that the general pattern
observed will not change. Despite these difficulties in assessing the occurrence of
the fungus on bats, our data are consistent with other studies (Puechmaille et al.,
2010; Martínková et al., 2010; Wibbelt et al., 2010), and also demonstrate that
the most commonly encountered bat species with G. destructans in Europe is the
largest species of Myotis on the continent, Myotis myotis. In countries/regions
(i.e., the Netherlands, Northwest Germany) where M. dasycneme is more com-
monly encountered in hibernacula, G. destructans prevalence can reach high lev-
els in that species. It is interesting to note that neither Pipistrellus pipistrellus nor
Miniopterus schreibersii have been observed with G. destructans (Puechmaille
et al., 2010; Martínková et al., 2010; Wibbelt et al., 2010), although these two
species are known to hibernate in aggregations of tens of thousands of individu-
als, especially the latter (Furman and Özgül, 2004; Nagy and Postwana, 2011;
Benda et al., 2003; Serra-Cobo et al., 1998). Although rare, hibernacula of a few
thousands and up to about 34,000 individuals are also known for species of Myo-
tis in Europe (Furman and Özgül, 2004; Nagy and Postwana, 2011; Kokurewicz,
2009; Arthur and Lemaire, 2009; Sachanowicz et al., 2006; Dietz et al., 2009).
APPENDIX A 395
During the two years monitoring at one site in Germany where G. destructans
prevalence reached high levels in March-April, not a single dead bat was found.
This is in agreement with previous studies (Puechmaille et al., 2010; Martínková
et al., 2010; Wibbelt et al., 2010) reporting that the presence of G. destructans
in bats from Europe is not associated with mass mortality. This sharply contrasts
with mass mortalities reported in North America where hundreds or thousands
of dead bats are found in hibernacula towards the end of the hibernation period.
Recent pathological investigations of bats dying from WNS in North America led
Cryan et al. (2010) to propose that mortality was caused by important disruptions
of wing-dependant physiological functions due to infection by G. destructans. In
North America, the fungus deeply invades wings tissues (Meteyer et al., 2009)
and causes damages that are thought to alter homeostasis and water balance,
resulting in more frequent arousals than bats can afford with their fat reserves,
leading to death by starvation (Cryan et al., 2010). The pathology associated with
G. destructans colonisation in Europe is not yet known. We believe that the first
step in understanding mortality differences between bats from Europe and North
America rely on understanding pathological differences incurred by the fungus
on the bats’ wings. As a result, we urge the necessity to carry out pathological
investigation of live bats from Europe colonised by G. destructans. Despite the
absence of mortality associated with the presence of G. destructans in Europe,
it would be necessary to investigate whether chronic infections with the fungus
are compromising the health of individuals, especially in M. Myotis and M.
dasycneme, which show high prevalence of the fungus towards the end of the
hibernation period.
Phylogeographic studies of European bat species have shown that in the
last 100,000 years, some species colonised Europe from Western Asia (Flanders
et al., 2009), including Myotis blythii (Berthier et al., 2006; Currat et al., 2008)
which has been found with G. destructans (Wibbelt et al., 2010). Assuming that
G. destructans can be transported over long distances by bats, we speculate that
the distribution of G. destructans is probably not limited to Europe and possibly
extends eastwards into Russia, Western and Central Asia. Further surveys are
necessary to clarify the global distribution of G. destructans.
Conclusions
We have shown here that G. destructans, the most likely causative agent of
WNS in North America, is widespread in Europe, but is not associated with mass
mortality. The prevalence of visible fungal growth on bats increases in February/
March before sharply decreasing when bats emerge from hibernation. We also
isolated viable G. destructans from the walls of an underground site suggesting
that the hibernacula could act as passive vectors and/ or reservoirs for G. de-
structans and therefore, might play an important role in the transmission process.
Further research is needed to clarify the global prevalence of G. destructans and
APPENDIX A 397
Sample collection
During ongoing population censuses carried out at hibernacula in different
countries across Europe and during additional hibernacula surveys carried out for
the purpose of this study, information on bats with visible white fungal growth
on snouts and/or ears was recorded. Whenever possible, sterile dry cotton swabs
(Puechmaille et al., 2010) or adhesive tape touch imprints (Wibbelt et al., 2010)
were used to collect fungal material from the bats. In Estonia, samples were
collected from the wall of the tunnel where a bat with characteristic white fun-
gus was observed nine days prior to the sampling. Where no sample collection
was possible, a photograph was taken of the bat (photographic record). In cases
where neither sample collection nor photographic evidence was obtained, the
record was classified as visual observation. Live hibernating bats with powdery,
white fungal growth on their noses were considered suspects of infection by G.
destructans (Gd-suspects) but not suspected of having WNS. There is presently
no data supporting the occurrence of WNS in Europe and the co-occurrence of
the fungus with lesions characteristic of WNS (Meteyer et al., 2009) has not (yet)
been reported in Europe (Wibbelt et al., 2010; Barlow et al., 2005). Although,
prevalence of G. destructans can reach high levels in some European species
(i.e., Myotis myotis, M. dasycneme) in late winter (especially in March), it can
be expected that by chance alone some bats dying from causes unrelated to the
presence of G. destructans will also be carrying the fungus. Unless the criteria
for the diagnosis of WNS are met (confirmation by histo-pathology and PCR)
(Meteyer et al., 2009) WNS should not be assumed as a cause of mortality in dead
bats found in hibernacula of Europe. Various species of fungi have been identified
on dead bats (Wibbelt et al., 2010; Voyron et al., 2011), most of them likely being
saprophytes that colonise bat carcasses post-mortem.
Fungal cultures
In the laboratory, samples were treated as in Puechemaille et al. (2010) for
swabs and following Wibbelt et al. (2010) for touch imprints. Briefly, swabs
were streak-plated onto plates of Sabouraud’s agar, supplemented with 0.1%
mycological peptone. For touch imprints, small areas with fungal conidia char-
acteristic of G. destructans were identified by light microscopy and the tape was
disinfected and excised before being transferred for culture to Sabouraud’s agar.
398 FUNGAL DISEASES
The plates were sealed with parafilm and incubated inverted in the dark at 10°C.
A fungal growth developed within 14 days, from which single spore cultures
were established.
Molecular identification
Each culture was sequenced for one molecular marker, the rRNA gene
internal transcribed spacer (ITS, ca. 930 bp.) region (ITS1, 5.8S, and ITS2) to
further confirm species identity. The DNA extraction, PCR amplification and
DNA sequencing followed protocols described in Puechmaille et al. (2010).
Briefly, DNA was extracted using the Qiagen Blood and Tissue kit following the
manufacturer’s instructions with slight modifications (after step 3, we added an
incubation time of 10 minutes at 70°C). PCR reactions were carried out in 25
mL containing 1 mL of DNA extract (at 10–75 ng/mL), 1.5 mmol/L MgCl2, 0.4
mmol/L each primer (Forward: ITS4, 5′-TCCTCCGCTTATTGATATGC – 3′;
Reverse: ITS5, 5′-GGAAGTAAAAGTCGTAACAAGG – 3′; (White et al., 1990),
0.2 mmol/L dNTP, 1x PCR buffer and 1 U Platinum Taq DNA Polymerase High
Fidelity (Invitrogen). PCR cycling conditions were: initial step 15′ at 95°C, then
10 cycles of 30″ at 95°C, 1′45″ at 60°C (reduce of 2°C every 2 cycles), 1′ at
72°C, following by 30 cycles of 30″ at 95°C, 1′45″ at 50°C and 1′ at 72°C. PCR
products were purified and sequenced by Macrogen Inc. (Seoul, Korea) in both
directions using the PCR primers. Complementary sequences were assembled
and edited for accuracy using CodonCode Aligner 3.0.3 (www.codoncode.com/
aligner/download.htm).
Supporting Information
FIGURE A21-S1╇ Monitoring of bats at an hibernaculum in Germany during (A) the win-
ter 2006/2007 (September 5th Figure
2006 untilA21-S1.eps
April 19th 2007) and (B), the winter 2007/2008
bitmap
(August 28th 2007 until April 23rd 2008). The blue line represents the total number of bats
counted whereas the green line represents the number of bats with visible white fungal
growth (Gd-suspects). Dotted vertical lines separate counts from each month. Note that
the number of counts per month was not equal between months. In (B), the black line rep-
resents the total number of bats counted whereas the blue line represents the total number
of bats bar one portion of the hibernaculum where bats grouped densely (ca. 20 individu-
als) and did not allow a reliable identification of the number of bats with white fungal
growth. The green line represents the number of bats with visible white fungal growth
(Gd-suspects) counted at the hibernaculum without considering individuals densely group-
ing at one place in the hibernaculum. The group of about 20 individuals formed while the
hibernaculum was partially flooded, likely as a result of bats changing position to avoid
drowning. Note that the right Y-axis scale is different between (A) and (B).
400 FUNGAL DISEASES
Acknowledgments
We would like to thank Dóczy Annamária, Andriy-Taras Bashta, Frédéric
Blanc, Sándor Boldogh, Gaby Bollen, Thomas Chatton, Emrah Coraman, Jére
Csaba, Simon Dutilleul, Mehmet Doker, Oliver Gebhardt, Lena Godlevska, René
Janssen, Daniel Lefèvre, Barti Levente, Vadim Martyniuk, Gerhard Mascher,
Mykola Matveev, Bernd Ohlendorf, Rian Pulles, Tony Rock, Wolfgang Rackow,
Sébastien Roué, Bücs Szilárd, Abigel Szodoray-Parádi, Farkas Szodoray-Parádi
and Julien Vittier for providing us with their field observations. The comments of
Paul Cryan, Paul Racey, Natalia Martínková and an anonymous reviewer helped
to improve a previous version of the manuscript.
Author Contributions
Conceived and designed the experiments: SJP GW VK ECT. Performed the
experiments: SJP GW HF VK KM AK. Analyzed the data: SJP GW. Contributed
reagents/materials/analysis tools: SJP GW VK HF KM AK FF WB CB TB TC
MD TG AJH FH GH MH CJ YLB LL MM BM KP MS AW UZ ECT. Wrote the
paper: SJP GW.
References
A national plan for assisting states, federal agencies, and tribes in managing white-nose syndrome in
bats. 2010 Draft v. 10.21.2010.16 p.
Arlettaz R, Ruchet C, Aeschimann J, Brun E, Genoud M, et al. (2000) Physiological traits affecting
the distribution and wintering strategy of the bat Tadarida teniotis. Ecology 81:1004–1014.
Arthur L, Lemaire M (2009) Les Chauves-souris de France, Belgique, Luxembourg et Suisse: Mèze:
Biotope, Paris: Muséum national d’Histoire naturelle. 576 p.
Barlow A, Ford S, Green R, Morris C, Reaney S (2009) Investigation into suspected white-nose
syndrome in two bat species in Somerset. Vet Rec 165:481–482.
Bastian F, Alabouette C, Saiz-Jimenez C (2009) The impact of arthropods on fungal community
structure in Lascaux Cave. J Appl Microbiol 106:1456–1462.
Battersby J (2010) Guidelines for surveillance and monitoring of European bats. Bonn, Germany.
95 p.
Benda P, Ivanova T, Horáček I, Hanák V, Cerveny J, et al. (2003) Bats (Mammalia: Chiroptera) of
the Eastern Mediterranean. Part 3. Review of bat distribution in Bulgaria. Acta Soc Zool Bohem
67:245–357.
Berthier P, Excoffier L, Ruedi M (2006) Recurrent replacement of mtDNA and cryptic hybridization
between two sibling bat species Myotis Myotis and Myotis blythii. Proc R Soc B 273:3101–3109.
Blehert DS, Hicks AC, Behr M, Meteyer CU, Berlowski-Zier BM, et al. (2009) Bat white-nose syn-
drome: an emerging fungal pathogen? Science 323: 227.
Brack V, Twente JW (1985) The duration of the period of hibernation of three species of vespertilionid
bats. I. Field studies. Can J Zoolog 63:2952–2954.
Cabrera A (1904) Ensayo monogra´fico sobre los quirópteros de España. Mem Soc Españ Hist Nat
2:249–292.
Chaturvedi V, Springer DJ, Behr MJ, Ramani R, Li X, et al. (2010) Morphological and molecular
characterizations of psychrophilic fungus Geomyces destructans from New York bats with White
Nose Syndrome (WNS). PLoS ONE 5: e10783.
APPENDIX A 401
Courtin F, Stone WB, Risatti G, Gilbert K, Van Kruiningen HJ (2010) Pathologic findings and liver
elements in hibernating bats with white-nose syndrome. Vet Pathol 47:214–219.
Cryan P, Meteyer CU, Boyles JG, Blehert DS (2010) Wing pathology of whitenose syndrome in bats
suggests life-threatening disruption of physiology. BMC Biol 8: 135.
Currat M, Ruedi M, Petit RJ, Excoffier L (2008) The hidden side of invasions: massive introgression
by local genes. Evolution 62:1908–1920.
Dietz C, Von Helversen O, Dietmar N (2009) Bats of Britain, Europe & Northwest Africa. London:
A & C Black Publishers Ltd. 400 p.
Flanders J, Jones G, Benda P, Dietz C, Zhang S, et al. (2009) Phylogeography of the greater horseshoe
bat, Rhinolophus ferrumequinum: contrasting results from mitochondrial and microsatellite data.
Mol Ecol 18:306–318.
Furman A, Özgül A (2004) The distribution of cave-dwelling bats and conservation status of under-
ground habitats in Northwestern Turkey. 120:243–248.
Garcia-Mudarra JL, Ibañez C, Juste J (2009) The straits of Gibraltar: barrier or bridge to Ibero-
Moroccan bat diversity? Biol J Linn Soc 96:434–450.
Gargas A, Trest MT, Christiensen M, Volk TJ, Blehert DS (2009) Geomyces destructans sp. nov. as-
sociated with bat white-nose syndrome. Mycotaxon 108:147–154.
Geomyces destructans Widespread in Europe White nose syndrome science strategy meeting II, 2009.
Consensus Statement. Austin, Texas.
Groth I, Vetermann R, Scuetze B, Schumann P, Saiz-Jimenez C (1999) Actinomycetes in karstic caves
of northern Spain (Altamira and Tito Bustillo). J Microbiol Meth 36:115–122.
Hallam TG, McCracken GF (2011) Management of the panzootic White-Nose Syndrome through
culling of bats. Conserv Biol 25:189–194.
Ibañez C, Garcia-Mudarra JL, Ruedi M, Stadelmann B, Juste J (2006) The Iberian contribution to
cryptic diversity in European bats. Acta Chiropt 8:277–297.
Kannan K, Hun Yun S, Rudd RJ, Behr M (2010) High concentrations of persistent organic pollutants
including PCBs, DDT, PBDEs and PFOS in little brown bats with white-nose syndrome in New
York, USA. Chemosphere 80:613–618.
Kokurewicz T (2009) Management Plan for the Natura 2000 site ‘‘Nietoperek’’ (Western Poland).
International Conference Military Heritage, Utrecht, The Netherlands.
KubátováA, Dvořák L (2005) Entomopathogenic fungi associated with insect hibernating in under-
ground shelters. Czech Mycol 57:221–237.
Larcher G, Bouchara JP, Pailley P, Montfort D, Beguin H, et al. (2003) Fungal biota associated with
bats in Western France. J Mycol Méd 13:29–34.
Lindner DL, Gargas A, Lorch JM, Banik MT, Glaser J, et al. (2010) DNA-based detection of the
fungal pathogen Geomyces destructans in soils from bat hibernacula. Mycologia. In press.
Martínková N, Bačkor P, Bartonička T, Blažkova´ P, Červený J, et al. (2010) Increasing incidence
of Geomyces destructans fungus in bats from the Czech Republic and Slovakia. PLoS ONE
5:e13853.
Mayer F, Dietz C, Kiefer A (2007) Molecular species identification boosts bat diversity. Front Zool
4:4.
Meteyer CU, Buckles EL, Blehert DS, Hicks AC, Green DE, et al. (2009) Histopathologic criteria to
confirm white-nose syndrome in bats. J Vet Diagn Invest 21:411–414.
Mosca AML, Campanino F (1962) Analisi micologiche del terreno di grotte piemontesi. Allionia.
pp. 27–43.
Nagy ZL, Postwana T (2011) Seasonal and geographical distribution of cavedwelling bats in Roma-
nia: implications for conservation. Anim Conserv 14:74–86.
Nováková A (2009) Microscopic fungi isolated from the Domica Cave system (Slovak Karst National
Park, Slovakia). A review. Int J Speleol 38:71–82.
Nováková A, Kolařik M (2010) Chrysosporium speluncarum, a new species resembling Ajellomyces
capsulatus, obtained from bat guano in caves of temperate Europe. Mycol Progress 9:253–260.
402 FUNGAL DISEASES
Park KJ, Jones G, Ransome RD (2000) Torpor, arousal and activity of hibernating Greater Horseshoe
bats (Rhinolophus ferrumequinum). Funct Ecol 14:580–588.
Parsons KN, Jones G (2003) Dispersion and habitat use by Myotis daubentonii and Myotis nattereri
during the swarming season: implications for conservation. Anim Conserv 6:283–290.
Parsons KN, Jones G, Davidson-Watts I, Greenaway F (2003a) Swarming of bats at underground
sites in Britain-implications for conservation. Biol Conserv 111:63–70. Geomyces destructans
Widespread in Europe PLoS ONE | www.plosone.org 10 April 2011 | Volume 6 | Issue 4 | e19167
Parsons KN, Jones G, Greenaway F (2003b) Swarming activity of temperate zone microchiropteran
bats: effect of season, time of night and weather conditions. J Zool 261:257–264.
Puechmaille SJ, Verdeyroux P, Fuller H, Ar Gouilh M, Bekaert M, et al. (2010) White-nose syndrome
fungus (Geomyces destructans) in bat, France. Emerging Infectious Diseases 16:290–293.
Puechmaille SJ, Fuller H, Teeling EC (2011) Effect of sample preservation methods on the viability
of Geomyces destructans, the fungus associated with whitenose syndrome in bats. Acta Chiropt.
In press.
Ransome RD (1971) The effect of ambient temperature on the arousal frequency of the hibernating
greater horseshoe bat, Rhinolophus ferrumequinum, in relation to site selection and the hiberna-
tion state. J Zool 164:353–371.
Reichard JD, Kunz TH (2009) White-nose syndrome inflicts lasting injuries to the wings of little
brown Myotis (Myotis lucifugus). Acta Chiropt 11:457–464.
Rice AV, Currah RS (2006) Two new species of Pseudogymnoascus with Geomyces anamorphs and
their phylogenetic relationship with Gymnostellatospora. Mycologia 98:307–318.
Rivers NM, Butlin RK, Altringham JD (2005) Genetic population structure of Natterer’s bats ex-
plained by mating at swarming sites and philopatry. Mol Ecol 14:4299–4312.
Rivers NM, Butlin RK, Altringham JD (2006) Autumn swarming behaviour of Natterer’s bats in the
UK: population size, catchment area and dispersal. Biol Conserv 127:215–226.
Rodrigues L, Zahn A, Rainho A, Palmeirim JM (2003) Contrasting the roosting behaviour and phe-
nology of an insectivorous bat (Myotis Myotis) in its southern and northern distribution ranges.
Mammalia 67:321–335.
Sachanowicz K, Sachanowicz M, Piksa K (2006) Distribution patterns, species richness and status of
bats in Poland. Vespertilio 9–10:151–173.
Senior P, Butlin RK, Altringham JD (2005) Sex and segregation in temperate bats. Proc R Soc Lond
B 272:2467–2473.
Serra-Cobo J, Sanz-Trullen V, Martinez-Rica JP (1998) Migratory movements of Miniopterus sch-
reibersii in the north-east of Spain. Acta Theriol 43:271–283.
Twente JW, Twente J, Brack V (1985) The duration of the period of hibernation of three species of
vespertilionid bats. II. Laboratory studies. Can J Zoolog 63:2955–2961.
Voyron S, Lazzari A, Riccucci M, Calvini M, Varese GC (2011) First mycological investigations on
Italian bats. Hystrix. In press.
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, Inc., New York. pp. 315–322.
Wibbelt G, Kurth A, Hellmann D, Weishaar M, Barlow A, et al. (2010) White-Nose Syndrome fungus
(Geomyces destructans) in bats, Europe. Emerging Infectious Diseases 16:1237–1242.
Appendix B
Agenda
8:45–9:30: The “Good,” the “Bad,” and the “Ugly”: Fungi Mold Your
World
Meredith Blackwell, Ph.D., Louisiana State University
403
404 FUNGAL DISEASES
10:15–10:45: Discussion
10:45–11:00: Break
Session I
Environmental Factors Influencing the Emergence
and Spread of Fungal Diseases
Moderator: Steven Brickner, Ph.D.
1:00–1:20: Discussion
1:20–2:00: Lunch
APPENDIX B 405
Session II
Case Studies of Emerging Fungal Diseases
of Humans, Animals, and Plants
Moderator: Erica Rosenblum, Ph.D.
3:30–3:45: Break
4:45–5:15: Discussion
Session III
Host and Pathogen Factors Influencing the
Emergence and Spread of Fungal Diseases
Moderator: Jeff Duchin, M.D.
10:15–10:30: Break
12:00–12:30: Discussion
12:30–1:15: Lunch
APPENDIX B 407
Session IV
Surveillance, Detection, and Response
Moderator: Steve Morse, Ph.D.
3:15–3:45: Break
Acronyms
409
410 FUNGAL DISEASES
TB tuberculosis
TGen Translational Genomics Research Institute
Glossary
Aggressive: In plant pathology, the quality of being able to cause more disease
more quickly on susceptible host plants.
Antibiotic: Class of substances that can kill or inhibit the growth of some groups
of microorganisms. Used in this report to refer to chemicals active against bacte-
ria. Originally antibiotics were derived from natural sources (e.g., penicillin from
molds), but many currently used antibiotics are semi-synthetic and modified with
additions of manmade chemical components.
Antifungal: Substances that can kill or inhibit the growth of fungal organisms.
413
414 FUNGAL DISEASES
Assembling the Tree of Life program: This National Science Foundation pro-
gram was created in 2004 with the goal of constructing the evolutionary history
for all major lineages of life. The program supports efforts to classify all major
groups of organisms and to reveal the pattern of historical relationships that
would explain similarities and differences among them.
The program has three goals: (1) Creation and support of multidisciplinary
teams of investigators to acquire and integrate molecular and morphological
evidence on both extant and extinct organisms in order to resolve phylogenetic
relationships of large, deep branches of the Tree of Life; (2) Research and de-
velopment of tools for computational phylogenetics and phyloinformatics to
improve assessment, predictive capabilities, and the visualization and navigation
of the hierarchical structure in the Tree of Life; and€(3) Outreach and education
in comparative phylogenetic biology and paleontology.
Biological invasion: The process by which species (or genetically distinct popu-
lations), with no historical record in an area, breach biogeographic barriers and
extend their range.
Conidia: Asexually produced fungal spore. Most conidia are dispersed by the
wind and can endure extremes of cold, heat, and dryness. When conditions are
favorable, they germinate and grow into hyphae.
Corn Belt: The area in the Midwestern United States—roughly covering west-
ern Indiana, Illinois, Iowa, Missouri, eastern Nebraska, and eastern Kansas—in
which corn (maize) and soybeans are the dominant crops.
Cultivar: A variety of a plant that has been created or selected intentionally and
maintained through cultivation.
Disease: A situation in which infection has elicited signs and symptoms in the
infected individual; the infection has become clinically apparent. Some exposures
to infectious disease-causing agents can also produce asymptomatic illnesses that
can be spread to others.
Endophytes: Fungi that live inside the plant tissue, but without causing any
obvious negative effects.
Etiology: Science and study of the causes of diseases and their mode of operation.
Eukaryotic organism: One of the three domains of life. The two other domains,
Bacteria and Archaea, are prokaryotes and lack several features characteristic of
eukaryotes (e.g., cells containing a nucleus surrounded by a membrane and whose
DNA is bound together by proteins [histones] into chromosomes). Animals,
plants, and fungi are all eukaryotic organisms.
Expression vectors: A plasmid that is used to introduce a specific gene into a tar-
get cell. Once the expression vector is inside the cell, the protein that is encoded
by the gene is produced by the cellular-transcription and translation machinery
ribosomal complexes. The plasmid is frequently engineered to contain regula-
tory sequences that act as enhancer and promoter regions and lead to efficient
transcription of the gene carried on the expression vector.
Extreme weather: Refers to weather phenomena that are at the extremes of the
historical distribution and are rare for a particular place and/or time, especially
418 FUNGAL DISEASES
Food security: The availability of food and one’s access to it. A household
is considered food secure when its occupants do not live in hunger or fear of
starvation.
Fungi/fungal/fungus: For the purposes of this publication, the terms fungi, fun-
gal, and fungus are used inclusively to describe all organisms traditionally studied
by mycologists—including species that are now excluded from Kingdom Fungi
(e.g., Phytophthora ramorum and Phytophthora infestans) or whose relationship
to the fungal kingdom has yet to be determined (e.g., the microsporidia Nosemus
spp.).
Genomics: The study of all the genes in a person, as well as interactions of those
genes with each other and with that person’s environment.
Genus: A group of species with similar characteristics that are closely related.
Hyphae: Slender tubes that develop from germinated spores and form the struc-
tural parts of the body of a fungus. A large mass of hyphae is known as a myce-
lium, which is the growing form of most fungi.
Incubation period: The time from the moment of inoculation (exposure to the
infecting organism) to the appearance of clinical manifestations of a particular
infectious disease.
Infection: The invasion of the body or a part of the body by a pathogenic agent,
such as a microorganism or virus. Under favorable conditions the agent develops
or multiplies; the results may produce injurious effects. Infection should not be
confused with disease.
Innate immune response: Immune response (of both vertebrates and inverte-
brates) to a pathogen that involves the preexisting defenses of the body, such as
barriers formed by skin and mucosa, antimicrobial molecules, and phagocytes.
Such a response is not specific for the pathogen.
Invasive species: Non-native plants and animals that, when introduced to new
environments, reproduce or spread so aggressively that they harm their adopted
ecosystems. Also called: exotic, alien, and non-indigenous species.
Mendelian: A single gene disorder caused by a defect in one particular gene, and
characterized by how they are passed down in families.
Mitigation: Initiatives that reduce the risk from natural and man-made hazards.
Mycelia, mycelium: The mass of fine branching tubes (known as hyphae) that
forms the main growing structure of a fungus.
One Health: Holistic approach to preventing epizootic disease and for maintain-
ing ecosystem integrity for the benefit of humans, their domesticated animals, and
the foundation biodiversity that supports all life.
Pandemic: Disease outbreak occurring over a wide geographic area and affecting
an exceptionally high proportion of the population.
Pathology: The branch of medicine concerned with disease, especially its struc-
ture and its functional effects on the body.
Propagules: Any of various structures that can give rise to a new individual or-
ganism. (For fungi, propagules include spores or encapsulated yeast cells.)
Public health: The art and science of dealing with the protection and improve-
ment of community health by organized community effort and including preven-
tive medicine and sanitary and social health.
Race (plant pathology): A subspecies group of pathogens that infect a given set
of plant varieties.
Reservoir: Any person, animal, arthropod, plant, soil, or substance (or combina-
tion of these) in which an infectious agent normally lives and multiplies, on which
it depends primarily for survival, and in which it reproduces itself in such manner
that it can be transmitted to a susceptible vector or host.
Somatic cells: The cells of the body, with the exception of the reproductive cells
(gametes).
Weather: The state of the atmosphere over a short period of time, measured
in terms of wind, temperature, humidity, atmospheric pressure, cloudiness, and
precipitation. The difference between weather and climate is a measure of time:
Climate is how the atmosphere “behaves” over relatively long periods of time.
Xylose: A white crystalline sugar extracted from wood, straw, and corn.
Yeast: Any of various one-celled fungi that reproduce by budding and can cause
the fermentation of carbohydrates, producing carbon dioxide and ethanol.
Zoonotic infection: Infection that causes disease in human populations but that
can be perpetuated solely in non-human host animals (e.g., bubonic plague); may
be enzootic or epizootic.
Appendix E
David A. Relman, M.D. (Chair), is the Thomas C. and Joan M. Merigan Pro�
fessor in the Departments of Medicine and of Microbiology and Immunology at
Stanford University, and Chief of Infectious Diseases at the VA Palo Alto Health
Care System in Palo Alto, California. He received an S.B. (biology) from Mas�
sachusetts Institute of Technology (1977), received his M.D. (magna cum laude)
from Harvard Medical School (1982), completed his clinical training in internal
medicine and infectious diseases at Massachusetts General Hospital, served as a
postdoctoral fellow in microbiology at Stanford University, and joined the faculty
at Stanford in 1994.
Dr. Relman’s current research focus is the human indigenous microbiota (mi-
crobiome) and, in particular, the nature and mechanisms of variation in pat�terns
of microbial diversity within the human body as a function of time (micro�bial
succession), space (biogeography within the host landscape), and in response to
perturbation, for example, antibiotics (community robustness and resilience). One
of the goals of this work is to define the role of the human microbiome in health
and disease. This research integrates theory and methods from ecology, popu�
lation biology, environmental microbiology, genomics, and clinical medicine.
During the past few decades, his research directions have also included patho�
gen discovery and the development of new strategies for identifying previously
unrecognized microbial agents of disease. This work helped to spearhead the
application of molecular methods to the diagnosis of infectious diseases in the
1990s. His research has emphasized the use of genomic approaches for exploring
host–microbe relationships. Past scientific achievements include the description
of a novel approach for identifying previously unknown pathogens; the identi�
fication of a number of new human microbial pathogens, including the agent of
427
428 FUNGAL DISEASES
Whipple’s disease; and some of the most extensive and revealing analyses to date
of the human indigenous microbial ecosystem.
Dr. Relman advises the U.S. government, as well as nongovernmental organi�
zations, in matters pertaining to microbiology, emerging infectious diseases, and
biosecurity. He is a member of the National Science Advisory Board for Biosecu�
rity, a member of the Physical and Life Sciences Directorate Review Commit-
tee for Lawrence Livermore National Laboratory, and he advises several U.S.
govern�ment departments and agencies on matters related to pathogen diversity,
the future life sciences landscape, and the nature of present and future biological
threats. He has served as Chair of the Board of Scientific Counselors of the Na-
tional Institute of Dental and Craniofacial Research (National Institutes of Health
[NIH]) and as a member of the Board of Directors, Infectious Diseases Society
of America (IDSA). Dr. Relman is currently vice-chair of a National Academy of
Sciences (NAS) study of the science underlying the Federal Bureau of Investiga-
tion inves�tigation of the 2001 anthrax mailings, and he cochaired a 3-year NAS
study that produced a widely cited report entitled Globalization, Biosecurity, and
the Future of the Life Sciences (2006). He is a Fellow of the American Academy
of Micro�biology and a member of the Association of American Physicians. Dr.
Relman received the Squibb Award from the IDSA in 2001 and was the recipient
of both the NIH Director’s Pioneer Award and the Distinguished Clinical Scientist
Award from the Doris Duke Charitable Foundation in 2006.
Lonnie J. King, D.V.M. (Vice-Chair), is the 10th dean of the College of Veteri-
nary Medicine at the Ohio State University (OSU). In addition to leading this
college, Dr. King is also a professor of preventive medicine and holds the Ruth
Stanton Endowed Chair in Veterinary Medicine. Before becoming dean at OSU,
he was the direcÂ�tor of CDC’s new National Center for Zoonotic, Vector-Borne,
and Enteric Dis�eases (NCZVED). In this new position, Dr. King leads the Cen-
ter’s activities for surveillance, diagnostics, disease investigations, epidemiol-
ogy, research, public education, policy development, and disease prevention and
control programs. NCZVED also focuses on waterborne, foodborne, vectorborne,
and zoonotic diseases of public health concern, which also include most of CDC’s
select and bioterrorism agents, neglected tropical diseases, and emerging zoono-
ses. Before serving as director, he was the first chief of the agency’s Office of
Strategy and Innovation.
Dr. King served as dean of the College of Veterinary Medicine, Michigan
State University, from 1996 to 2006. As at OSU, he served as the CEO for aca�
demic programs, research, the teaching hospital, the diagnostic center for popula-
tion and animal health, basic and clinical science departments, and the outreach
and continuing education programs. As dean and professor of large-animal clini-
cal sciences, Dr. King was instrumental in obtaining funds for the construction
of a $60 million Diagnostic Center for Population and Animal Health; he initi-
ated the Center for Emerging Infectious Diseases in the college, he served as the
campus leader in food safety, and he had oversight for the National Food Safety
and Toxicology Center.
In 1992, Dr. King was appointed administrator for the Animal and Plant
Health Inspection Service (APHIS), U.S. Department of Agriculture (USDA),
in Washington, DC. In this role, he provided executive leadership and direction
for ensuring the health and care of animals and plants, to improve agricultural
productivity and competitiveness, and to contribute to the national economy and
public health. Dr. King also served as the country’s chief veterinary officer for 5
years, worked extensively in global trade agreements within the North American
Free Trade Agreement and the World Trade Organization, and worked extensively
with the World Animal Health Asso�ciation. During this time he was the Deputy
430 FUNGAL DISEASES
Ruth L. Berkelman, M.D., is the Rollins Professor and director of the Center for
Public Health Preparedness and Research at the Rollins School of Public Health,
Emory University, in Atlanta. She received her A.B. from Princeton University
and her M.D. from Harvard Medical School. Board certified in pediatrics and
internal medicine, she began her career at CDC in 1980 and later became deputy
director of NCID. She also served as a senior advisor to the director of CDC and
as assistant surgeon general in the U.S. Public Health Service. In 2001 she came
to her current position at Emory University, directing a center focused on emerg�
ing infectious diseases and other urgent threats to health, including terrorism. She
has also consulted with the biologic program of the Nuclear Threat Initiative and
is most recognized for her work in infectious diseases and disease surveillance.
She was elected to the IOM in 2004. Currently a member of the Board on Life
Sciences of the National Academies, she also chairs the Board of Public and
Scientific Affairs at the ASM.
his main commitment was to the Defense Threat Reduction Agency’s Global
Bioengagement Program.
Paula R. Bryant, Ph.D., is Chief of the Medical S&T Division, Chemical and
Biological Defense Program at the Defense Threat Reduction Agency (DTRA) in
Fort Belvoir, Virginia. As the Chief of the Medical S&T Division, Bryant inter-
faces with all levels of the Department of Defense and DTRA to plan, coordinate,
integrate, and execute program activities necessary to provide timely and effective
medical countermeasures against Chemical, Biological and Radiological (CBR)
threats to U.S. interests worldwide. She also served as a Senior Scientist and Se-
nior S&T Manager while at DTRA. Prior to joining DTRA, she was an Assistant
Professor in the Department of Microbiology at The Ohio State University. She
received her Ph.D. in Microbiology and Immunology from the Baylor College
of Medicine.
Arturo Casadevall, M.D., Ph.D.,3 is the Leo and Julia Forchheimer Professor
of Microbiology & Immunology at the Albert Einstein College of Medicine of
Yeshiva University in the Bronx, New York. He is Chairman of the Department of
Microbiology and Immunology and served as Director of the Division of Infec-
tious Diseases at the Montefiore Medical Center at the Albert Einstein College
of Medicine from 2000–2006. Dr Casadevall received both his M.D. and Ph.D.
(biochemistry) degrees from New York University in New York, New York. Sub-
sequently, he completed internship and residency in internal medicine at Bellevue
Hospital in New York, New York. Later he completed subspecialty training in
Infectious Diseases at the Montefiore Medical Center and Albert Einstein College
of Medicine. Dr. Casadevall major research interests are in fungal pathogenesis
and the mechanism of antibody action. In the area of Biodefense Dr. Casade-
vall has an active research program to understand the mechanisms of antibody-
mediated neutralization of Bacillus anthracis toxins. He has authored over 500
scientific papers. Dr. Casadevall was elected to membership in the American
Society for Clinical Investigation, the American Academy of Physicians, and the
American Academy of Microbiology. He was elected a fellow of the American
Academy for the Advancement of Science and has received numerous honors
including the Solomon A Berson Medical Alumni Achievement Award in Basic
Science from the NYU School of Medicine, the Maxwell L. Littman Award
(mycology award), the Rhoda Benham Award from Medical Mycology Society
of America, and the Kass Lecture of the Infectious Disease Society of America.
Dr. Casadevall is the Editor in Chief of mBio, the first open access general jour-
nal of the American Society of Microbiology. He serves in the editorial board of
the Journal of Clinical Investigation, The Journal of Experimental Medicine and
The Journal of Infectious Diseases. Previously he served as Editor of Infection
and Immunity. He has served in numerous NIH committees including those that
drafted the NIAID Strategic Plan and the Blue Ribbon Panel on Biodefense Re-
search. Dr. Casadevall served on the NAS committee that reviewed the science
behind the FBI investigation of the anthrax attacks in 2001. He is currently a
member of the National Science Advisory Board for Biosecurity and co-chaired
the NIAID Board of Scientific counselors.
Since he joined the Einstein faculty in 1992 Dr. Casadevall has mentored
dozens of graduate students, postdoctoral fellows, and junior faculty. Many of
his trainees have gone on to have highly successful careers in science and several
have currently AECOM faculty. From 2000–2006 Dr. Casadevall was director
of the Division of Infectious Diseases at AECOM-Montefiore and oversaw the
expansion of its research program. In 2001 Dr. Casadevall received the Samuel
M. Rosen outstanding teacher award and in 2008 he was recognized the American
Society of Microbiology with the William Hinton Award for mentoring scientists
from underrepresented groups.
Dr. Duchin’s peer review publications and research interests focus on communi-
cable diseases of public health significance, and he has authored text book chap-
ters on the epidemiology of HIV/AIDS, bioterrorism, and outbreak investigations.
Mark B. Feinberg, M.D., Ph.D., is vice president for medical affairs and policy
in global vaccine and infectious diseases at Merck & Co., Inc., and is responsible
for global efforts to implement vaccines to achieve the greatest health benefits,
including efforts to expand access to new vaccines in the developing world. Dr.
Feinberg received a bachelor’s degree magna cum laude from the University of
Pennsylvania in 1978 and his M.D. and Ph.D. degrees from Stanford University
School of Medicine in 1987. His Ph.D. research at Stanford was supervised by
Dr. Irving Weissman and included time spent studying the molecular biology of
the human retroviruses—human T-cell lymphotrophic virus, type I (HTLV-I) and
HIV—as a visiting scientist in the laboratory of Dr. Robert Gallo at the National
438 FUNGAL DISEASES
Cancer Institute. From 1985 to 1986, Dr. Feinberg served as a project officer for
the IOM Committee on a National Strategy for AIDS. After receiving his M.D.
and Ph.D. degrees, Dr. Feinberg pursued postgraduate residency training in inter-
nal medicine at the Brigham and Women’s Hospital of Harvard MediÂ�cal School
and postdoctoral fellowship research in the laboratory of Dr. David Baltimore
at the Whitehead Institute for Biomedical Research. From 1991 to 1995, Dr.
Feinberg was an assistant professor of medicine and microbiology and immunol-
ogy at the University of California, San Francisco (UCSF), where he also served
as an attending physician in the AIDS-oncology division and as director of the
virology research laboratory at San Francisco General Hospital. From 1995 to
1997, Dr. Feinberg was a medical officer in the Office of AIDS Research in the
Office of the Director of the NIH, the chair of the NIH Coordinating Committee
on AIDS Etiology and Pathogenesis Research, and an attending physician at the
NIH Clinical Center. During this period, he also served as executive secretary of
the NIH Panel to Define Principles of Therapy of HIV Infection. Prior to joining
Merck in 2004, Dr. Feinberg served as professor of medicine and microbiology
and immunology at the Emory University School of Medicine, as an investigator
at the Emory Vaccine Center, and as an attending physician at Grady Memorial
Hospital. At UCSF and Emory, Dr. Feinberg and colleagues were engaged in
the preclinical development and evaluation of novel vaccines for HIV and other
infectious diseases and in basic research studies focused on revealing fundamen�
tal aspects of the pathogenesis of AIDS. Dr. Feinberg also founded and served as
the medical director of the Hope Clinic of the Emory Vaccine Center—a cliniÂ�
cal research facility devoted to the clinical evaluation of novel vaccines and to
translational research studies of human immune system biology. In addition to
his other professional roles, Dr. Feinberg has also served as a consultant to, and
a member of, several IOM and NAS committees. Dr. Feinberg currently serves
as a member of the National Vaccine Advisory Committee and is a member of
the Board of Trustees of the National Foundation for Infectious Diseases. Dr.
Feinberg has earned board certification in internal medicine; he is a fellow of
the American College of Physicians, a member of the Association of American
Physicians, and the recipient of an Elizabeth Glaser Scientist Award from the
Pediatric AIDS Foundation and an Innovation in Clinical Research Award from
the Doris Duke Charitable Foundation.
land Security Under Secretary’s Award for Science and Technology. She is the
author of numerous journal articles and has presented her research at national
and international meetings.
Jesse L. Goodman, M.D., M.P.H., became chief scientist and deputy commis�
sioner for science and public health of FDA in 2009. He has broad responsibility
for and engagement in leadership and coordination of FDA’s crosscutting sciÂ�
entific and public health efforts. From 2003 to 2009, Dr. Goodman was director
of FDA’s Center for Biologics Evaluation and Research, which oversees mediÂ�cal
and public health activities critical to U.S. and global preparedness and the de-
velopment, evaluation, safety, quality, and availability of biologics. A graduate
of Harvard, Dr. Goodman received his M.D. from the Albert Einstein College
of Medicine and did residency and fellowship training at the Hospital of the
Univer�sity of Pennsylvania and at the University of California, Los Angeles
(UCLA), where he was also Chief Medical Resident. Prior to joining FDA, he
was professor of medicine and chief of infectious diseases at the University of
Minnesota, where he directed the multihospital infectious diseases research, train-
ing, and clinical programs, and where his NIH-funded laboratory first isolated and
characterized Anaplasma phagocytophilum, the infectious agent causing a new
tickborne disease, human granulocytic ehrlichiosis. Dr. Goodman has authored
numerous scientific papers and edited the book Tick Borne Diseases of Humans
(ASM Press, 2005). Dr. Goodman has been elected to the American Society for
Clinical Investigation and to the IOM of the NAS, where he is a longstanding
member of the Forum on Microbial Threats. He is an active clinician and teacher
who is board certified in internal medicine, oncology, and infectious diseases
and is staff physician and infectious diseases consultant at the National Naval
and Walter Reed Army Medical Centers. Dr. Goodman is adjunct professor of
medicine at the University of Minnesota.
Hygiene, and the 2007 Heinz Award on the Human Condition. In 2009 he was
appointed an honorary Commander of the Most Excellent Order of the British
Empire for services to global public health, and he was recently elected a Fel-
low of the Academy of Medical Sciences in the United Kingdom. Dr. Heymann
has been visiting professor at Stanford University, the University of Southern
California, and the George Washington University School of Public Health; has
published over 145 scientific articles on infectious diseases and related issues in
peer-reviewed medical and scientific journals; and has authored several chapters
on infectious diseases in medical textbooks. He is currently the editor of the 19th
edition of the Control of Communicable Diseases Manual, a joint publication of
the American Public Health Association and WHO.
Stephen Albert Johnston, Ph.D., is currently director of the Center for Inno-
vations in Medicine in the Biodesign Institute at Arizona State University. His
cen�ter focuses on formulating and implementing disruptive technologies for
basic problems in health care. The center has three divisions: Genomes to Vac-
cines, Cancer Eradication, and DocInBox. Genomes to Vaccines has developed
high-throughput systems to screen for vaccine candidates and is applying them to
predict and produce chemical vaccines. The Cancer Eradication group is working
on formulating a universal prophylactic vaccine for cancer. DocInBox is devel�
oping technologies to facilitate presymptomatic diagnosis. Dr. Johnston founded
the Center for Biomedical Inventions (also known as the Center for Translation
Research) at the University of Texas, Southwestern, the first center of its kind in
the medical arena. He and his colleagues have developed numerous inventions
and innovations, including the gene gun, genetic immunization, the tobacco
etch virus protease system, organelle transformation, digital optical chemistry
arrays, expression library immunization, linear expression elements, synbodies,
immunosignaturing diagnosis, and others. He also was involved in transcription
research for years, first cloning Gal4 and later discovering functional domains in
transcription factors and the connection of the proteasome to transcription. He has
been professor at the University of Texas Southwestern Medical Center at Dallas
and associate and assistant professor at Duke University. He has been involved
in several capacities as an adviser on biosecurity since 1996 and is a founding
member of BioChem 20/20.
Kent Kester, M.D., is currently the commander of the Walter Reed Army Insti�
tute of Research (WRAIR) in Silver Spring, Maryland. Dr. Kester holds an un-
dergraduate biology degree from Bucknell University (1982) and an M.D. from
Jefferson Medical College (1986). He completed his internship and residency in
internal medicine at the University of Maryland Hospital/Baltimore VA Medical
Center (1989) and a fellowship in infectious diseases at the Walter Reed Army
Medical Center (1995). A malaria vaccine researcher with over 50 authored or
coauthored scientific manuscripts and book chapters, Dr. Kester has played a
major role in the development of the candidate falciparum malaria vaccine known
as RTS,S, having safely conducted the largest number of experimental malaria
challenge studies ever attempted to date. Dr. Kester’s previous military medical
research assignments have included director of the WRAIR Malaria Serology
Reference Laboratory; chief, Clinical Malaria Vaccine Development Program;
chief of the WRAIR Clinical Trials Center; and director of the WRAIR Division
of Regulated Activities. He currently is a member of the Steering Committee
of the NIAID/Uniformed Services University of the Health Sciences Infectious
Disease Clinical Research Program, as well as multiple NIAID Safety Monitor�
ing Committees. He also serves as the consultant to the U.S. Army Surgeon
General in Medical Research and Development. Board certified in both internal
medicine and infectious diseases, Dr. Kester is also a fellow of both the Ameri�
APPENDIX E 445
can College of Physicians and the IDSA. He holds faculty appointments at both
the Uniformed Services University of the Health Sciences and the University of
Maryland School of Medicine.
Gerald T. Keusch, M.D., is associate provost and associate dean for global
health at Boston University and Boston University School of Public Health. He
is a graduate of Columbia College (1958) and Harvard Medical School (1963).
After completing a residency in internal medicine, fellowship training in infec�
tious diseases, and 2 years as an NIH research associate at the Southeast Asia
Treaty Organization Medical Research Laboratory in Bangkok, Thailand, Dr.
Keusch joined the faculty of the Mt. Sinai School of Medicine in 1970, where
he established a laboratory to study the pathogenesis of bacillary dysentery and
the biology and biochemistry of Shiga toxin. In 1979 he moved to Tufts Medi�
cal School and New England Medical Center in Boston to found the Division of
Geographic Medicine, which focused on the molecular and cellular biology of
tropical infectious diseases. In 1986 he integrated the clinical infectious diseases
program into the Division of Geographic Medicine and Infectious Diseases,
continuing as division chief until 1998. He has worked in the laboratory and
in the field in Latin America, Africa, and Asia on basic and clinical infectious
diseases and HIV/AIDS research. From 1998 to 2003, he was associate director
for international research and director of the Fogarty International Center at NIH.
Dr. Keusch is a member of the American Society for Clinical Investigation, the
Association of American Physicians, the ASM, and the IDSA. He has received
the Squibb (1981), Finland (1997), and Bristol (2002) awards of the IDSA. In
2002 he was elected to the IOM.
well as the 2001 anthrax attacks. She is a fellow of the IDSA and member of the
American Epidemiologic Society, the ASM, and the Council of State and Territo-
rial Epidemiologists. She served on IDSA’s Annual Meeting Scientific Program
Committee and currently serves on the society’s National and Global Public
Health Committee. In addition to her CDC position, she serves as clinical as-
sociate professor of medicine (infectious diseases) at Emory University. She is a
graduate of the National Preparedness Leadership Initiative at Harvard Uni�versity
and of the Public Health Leadership Institute at the University of North Carolina.
NIAID’s bilateral program with India, the Indo–U.S. Vaccine Action Program,
and he represents NIAID in the HHS Biotechnology Engagement Program with
Russia and related countries. He is a member of AAAS, the ASM, and the Na-
tional Asso�ciation of Science Writers. He is the author of numerous journal and
freelance articles.
John C. Pottage, Jr., M.D., has been vice president for Global Clinical Develop�
ment in the Infectious Disease Medicine Development Center at GlaxoSmith-
Kline since 2007. Previously he was senior vice president and chief medical
officer at Achillion Pharmaceuticals in New Haven, Connecticut. Achillion is
a small biotechnology company devoted to the discovery and development of
medicines for HIV, hepatitis C virus, and resistant antibiotics. Dr. Pottage ini-
tially joined Achillion in May 2002. Prior to Achillion, Dr. Pottage was medical
director of Antivirals at Vertex Pharmaceuticals. During this time he also served
as an asso�ciate attending physician at the Tufts New England Medical Center in
Boston. From 1984 to 1998, Dr. Pottage was a faculty member at Rush Medical
College in Chicago, where he held the position of associate professor, and also
served as the medical director of the Outpatient HIV Clinic at Rush-Presbyterian-
St. Luke’s Medical Center. While at Rush, Dr. Pottage was the recipient of several
teaching awards and is a member of the Mark Lepper Society. Dr. Pottage is a
graduate of St. Louis University School of Medicine and Colgate University.
David Rizzo7 received his Ph.D. in plant pathology from the University of Min-
nesota and joined the faculty of the University of California-Davis, Department
of Plant Pathology and the Graduate Group in Ecology in 1995. Research in his
lab focuses on the ecology and management of forest€tree diseases, including dis-
eases caused by both native and introduced pathogens. Research in the lab takes
a multiscale approach ranging from experimental studies on the basic biology of
organisms to field studies across forest landscapes. Active collaborations include
projects with landscape ecologists, epidemiologists, molecular biologists, ento-
mologists, and forest managers. The primary research effort in the lab is currently
Phytophthora€species in California coastal forests, with an emphasis on€Sudden
Oak Death. As part of his research on Sudden oak Death, Dr. Rizzo also serves
as the scientific advisor for the California Oak Mortality Task Force.In conifer
forests of the Sierra Nevada Mountains, the lab studies a variety of diseases and
their relationship to past and present forest management and conservation issues.
In addition to research, Dr. Rizzo teaches undergraduate and graduate courses in
mycology as well as introductory biology. Since 2004, he has been director of the
Science and Society program in the College of Agricultural and Environmental
Sciences. Science and Society is an academic program designed to offer students
the opportunity to discover the interdisciplinary connections that link the biologi-
cal, physical and social sciences with societal issues and cultural discourses.
Gary A. Roselle, M.D., is program director for infectious diseases for the VA
Central Office in Washington, DC, as well as the chief of the medical service at
the Cincinnati VA Medical Center. He is a professor of medicine in the Depart�
ment of Internal Medicine, Division of Infectious Diseases, at the University of
Cincinnati College Of Medicine. Dr. Roselle serves on several national advi�
sory committees. In addition, he is currently heading the Emerging Pathogens
Initiative for the VA. He has received commendations from the undersecretary
for health for the VA and the secretary of VA for his work in the Infectious Dis�
eases Program for the VA. He has been an invited speaker at several national
and international meetings and has published more than 90 papers and several
book chapters. Dr. Roselle received his medical degree from the OSU School of
Medicine in 1973. He served his residency at the Northwestern University School
of Medicine and his infectious diseases fellowship at the University of Cincinnati
School of Medicine.
Alan S. Rudolph, Ph.D., M.B.A., has led an active career in translating inter-
disciplinary life sciences into useful applications for biotechnology devel�opment.
His experience spans basic research to advanced development in academia, gov-
ernment laboratories, and most recently in the nonprofit and private sectors. He
has published more than 100 technical publications in areas including molecular
biophysics, lipid self-assembly, drug delivery, blood sub�stitutes, medical imag-
ing, tissue engineering, neuroscience, and diagnostics. As a National Research
Council Post-Doctoral Fellow, his earliest work at the U.S. Naval Research
Laboratory (NRL) demonstrated the translational value of strategies used by or-
APPENDIX E 451
In 2001, he moved back to the United States and became the director of the Re-
spiratory Disease Laboratory at the Naval Health Research Center in San Diego,
California. Febrile respiratory illness surveillance in recruits of all services was
expanded into shipboard populations, Mexican border populations, support for
outbreaks, and deployed settings. Validation and integration of new and emerging
advanced diagnostic capabilities, utilizing the archives of specimens maintained
at the laboratory, became a priority. A BSL-3-Enhanced was constructed. Projects
expanded in 2006 to clinical trials support as Dr. Russell became the Principal
Investiga�tor for the Navy site in the FDA Phase III adenovirus vaccines trial, and
more recently to support the Phase IV post-marketing trial of the recently FDA-
approved ACAM2000 smallpox vaccine.
Janet Shoemaker is director of the ASM’s Public Affairs Office, a position she
has held since 1989. She is responsible for managing the legislative and regula�
tory affairs of this 42,000-member organization, the largest single biological sci�
ence society in the world. Previously, she held positions as assistant director of
public affairs for the ASM; as ASM coordinator of the U.S.–U.S.S.R. Exchange
Program in Microbiology, a program sponsored and coordinated by the NSF and
the U.S. Department of State; and as a freelance editor and writer. She received
her baccalaureate, cum laude, from the University of Massachusetts and is a
graduate of the George Washington University programs in public policy and in
editing and publications. She is a member of Women in Government Relations,
the American Society of Association Executives, and AAAS. She has coauthored
articles on research funding, biotechnology, biodefense, and public policy issues
related to microbiology.
Terence Taylor is the founding president of the International Council for the
Life Sciences (ICLS). The ICLS is an independent nonprofit organization reg�
istered in the United States and in the European Union. The ICLS is designed
to promote best practices and codes of conduct for safety and security in rela�
tion to biological risks. Terence Taylor also served as the vice president, Global
Health and Security, at the Nuclear Threat Initiative. Prior to these appointments
Terence Taylor was assistant director at the International Institute for Strategic
Studies (IISS) in London and was president and executive director of IISS-US in
Washington, DC. At IISS, in addition to his overall program responsibilities, he
led the Institute’s work on life sciences and security. He has substantial experiÂ�
ence in international security policy matters as a U.K. government official (both
military and diplomatic) and for the United Nations (UN) both in the field and
at UN Headquarters. He was a commissioner and one of the Chief Inspectors
with the UN Special Commission on Iraq, with particular responsibilities for
biologi�cal issues. His government experience is related to both military field
operations and to the development and implementation of policies in relation to
arms control and nonproliferation treaties and agreements for both conventional
weapons and weapons of mass destruction and the law of armed conflict aspects
of International Human�itarian Law. He has also conducted consulting work on
political risk assessment and studies of the private biotechnology industry. He
was a Science Fellow at Stanford University’s Center for International Security
and Cooperation. He was an officer in the British Army with experience in many
parts of the world includ�ing UN peacekeeping, counterinsurgency, and counter-
terrorism operations.
Speaker Biographies
Karen Bartlett, Ph.D., M.Sc., received her B.A. from the University of Victo-
ria, her M.Sc. in Occupational Hygiene from the University of British Columbia
(UBC), and her Ph.D. in Interdisciplinary Studies (Environmental Health) also
from UBC. She completed postdoctoral training in Inhalation Toxicology at the
University of Iowa. She is an associate professor in the School of Environmental
Health, in the College of Interdisciplinary Studies at UBC. Dr. Bartlett’s current
research interests are in four thematic areas: environmental sources of infectious
disease; mold and building material interactions (the built environment); animal
models of lung infections for therapeutic protocols; and occupational and envi-
ronmental exposure to airborne biologic particles. Examples of recent research
are bioaerosol exposures in the built environment (including First Nations hous-
ing); exposures to compost workers; and environmental sources of Cryptococcus
gattii. Dr. Bartlett is an adjunct faculty in the Department of Pathology, UBC,
and an associate faculty member in the Institute of Resource and Environmental
Sustainability and the School of Populations and Public Health.
455
456 FUNGAL DISEASES
Dr. Blackwell, who is Boyd Professor at Louisiana State University, has been
involved in several projects involving the fungal systematics community. The
recent Deep Hypha and Assembling the Fungal Tree of Life projects involved
more than 100 mycologists from 25 countries in phylogenetic studies of major
groups of fungi and a phylogenetic classification used in many major publica-
tions. Dr. Blackwell is a Distinguished Mycologist of the Mycological Society of
America and Fellow of the British Mycological Society. She served as president
of the International Mycological Association and the Mycological Society of
America and is coauthor of Introductory Mycology, a widely used textbook cur-
rently under revision.
David Blehert, Ph.D., earned his doctorate degree in bacteriology in 1999 from
the University of Wisconsin–Madison where he studied the biotransformation
of munitions manufacturing wastes as mediated by soil bacteria. He then com-
pleted a postdoctoral fellowship at the National Institutes of Health in Bethesda,
Maryland where he investigated bacterial communication mechanisms among
constituents of the human dental plaque community. His research emphasis was
on the role of the signaling molecule autoinducer-2 in the formation of bacterial
biofilms. Dr. Blehert joined the U.S. Geological Survey–National Wildlife Health
Center in Madison, Wisconsin, in 2003 as the head of the diagnostic microbiology
laboratory. Current research projects under way in his laboratory include charac-
terization of microbial aspects of the pathogenesis and epidemiology of bat white-
nose syndrome; the use of molecular markers to understand the epidemiology of
avian cholera in wild waterfowl; and development of rapid in vitro techniques
for the detection of botulinum neurotoxins. His laboratory’s collaborative efforts
to identify the fungus that causes the skin infection hallmark of bat white-nose
syndrome were published in the January 9, 2009, issue of Science.
Arturo Casadevall, M.D., Ph.D., is the Leo and Julia Forchheimer Profes-
sor and Chair of Microbiology & Immunology at the Albert Einstein College
of Medicine. Dr. Casadevall received both his M.D. and Ph.D. (biochemistry)
degrees from New York University in New York. Subsequently, he completed
internship and residency in internal medicine at Bellevue Hospital in New York.
Later he completed subspecialty training in infectious diseases. Dr. Casadevall’s
major research interests are in fungal pathogenesis and the mechanism of anti-
body action. Dr. Casadevall has authored more than 470 scientific papers. He has
been elected to membership in the American Society for Clinical Investigation,
American Academy of Physicians, and American Academy of Microbiology. He
was elected a Fellow of the American Academy for the Advancement of Science
(AAAS) and has received numerous honors. In 2005, he received the Alumni
Award in basic science from his alma mater, New York University. Dr. Casadevall
is editor in chief of mBio and serves on the editorial board of the Journal of Clini-
APPENDIX F 457
John N. Galgiani, M.D., was born in San Francisco, received his B.A. from
Stanford University, his M.D. from Northwestern University, and a fellowship in
Infectious Diseases from Stanford. In 1978, Dr. Galgiani joined the faculty of the
University of Arizona currently he is Professor of Medicine. Dr. Galgiani has fo-
cused his career primarily on the special problems of coccidioidomycosis (Valley
Fever) and its impact on the general population and special groups such as organ
transplant recipients and patients with AIDS. For 19 years, he was project direc-
tor of an NIH-sponsored coccidioidomycosis clinical trials group. Dr. Galgiani’s
laboratory has collaborated in efforts to develop vaccines to prevent Valley Fever.
For the past 5 years, Dr. Galgiani has led a development program for nikkomycin
Z, a possible cure for Valley Fever, now in clinical trials. In 1996, Dr. Galgiani
founded the Valley Fever Center for Excellence to disseminate information about
Valley Fever, help patients with the severest complications of this disease, and
encourage research into the biology and diseases of its etiologic agent.
Julie Harris, Ph.D., M.P.H., received her bachelors’ degree from Rensselaer
Polytechnic Institute, her Ph.D. in microbiology from Columbia University and
her M.P.H. from Johns Hopkins in epidemiology before joining the CDC’s Epi-
demic Intelligence Service in 2007, where she worked on the prevention and
control of enteric infections in the United States and in several countries in Africa.
APPENDIX F 459
She joined the Mycotic Diseases Branch at the CDC in 2009, where her first task
was to create a surveillance system for Cryptococcus gattii, an emerging fungal
infection in the Pacific Northwest. She also works on mycotoxins in Bangladesh
and Guatemala, cryptococcal infections in Thailand, and coccidioidomycosis in
the United States.
Joseph Heitman, M.D., Ph.D., is chair and James B. Duke Professor, Depart-
ment of Molecular Genetics and Microbiology, Duke University. He received his
B.S. and M.S. in chemistry and biochemistry at the University of Chicago, and
M.D. and Ph.D. from Cornell and Rockefeller Universities. He was an EMBO
fellow at the Biocenter in Switzerland where he pioneered yeast as a model to
study immunosuppressive drugs. He elucidated the role of FKBP12 in forming
complexes with FK506 and rapamycin that inhibit cell signaling and growth
and discovered the targets of rapamycin TOR1/TOR2, pathways conserved from
yeasts to humans. Dr. Heitman has been at Duke since 1992, and focuses on
model and pathogenic fungi, studying the evolution of sex and roles of sexual
reproduction in microbial pathogens; how cells sense and respond to nutrients
and the environment; the targets and mechanisms of action of immunosuppres-
sive and antimicrobial drugs; and the genetic and molecular basis of microbial
pathogenesis and development. Their discovery of fungal unisexual mating has
implications for how sex might create diversity de novo with implications for
pathogen evolution and emergence.
Dr. Heitman received the Burroughs Wellcome Scholar Award in Molecular
Pathogenic Mycology, the ASBMB AMGEN award, and the Infectious Diseases
Society of America (IDSA) Squibb Award. He has taught the Molecular Mycol-
ogy Course at the MBL since 1998. He is an editor for Eukaryotic Cell, Fungal
Genetics and Biology, and PLoS Pathogens; a board member for PLoS Biology,
Current Biology, and Cell Host & Microbe; an advisory board member for the
Broad Institute Fungal Genome Initiative and the Department of Energy/JGI Fun-
gal Kingdom project; and cochair/chair for the FASEB Microbial Pathogenesis
conference (2011, 2013). He is a Fellow of the American Society for Clinical
Investigation, the IDSA, the American Academy of Microbiology, AAAS, and
the Association of American Physicians.
Steven Holland, M.D., received his M.D. and Medicine and Infectious Disease
training from Johns Hopkins. He came to NIH in 1989 as a National Research
Council fellow in the Laboratory of Molecular Microbiology, working on tran-
scriptional regulation of HIV. In 1991, Dr. Holland joined the Laboratory of
Host Defenses, shifting his research to the host side, with a focus on phagocyte
defects and their associated infections. His work centered on the pathogenesis
and management of chronic granulomatous disease, as well as other congenital
immune defects affecting phagocytes, including those predisposing to mycobac-
terial and fungal diseases. In 2004, he became chief of the Laboratory of Clinical
460 FUNGAL DISEASES
Luis Padilla, D.V.M., has served as the staff clinical veterinarian for the Smith-
sonian Conservation Biology Institute of the National Zoological Park in Front
Royal, Virginia, since 2007. He received his Doctorate of Veterinary Medicine
from Cornell University, where he also obtained a B.S. degree in biology. Dr.
462 FUNGAL DISEASES
David Rizzo, Ph.D., received his Ph.D. in plant pathology from the University of
Minnesota and joined the faculty of the University of California-Davis, Depart-
ment of Plant Pathology and the Graduate Group in Ecology in 1995. Research
in his lab focuses on the ecology and management of forest tree diseases, includ-
ing diseases caused by both native and introduced pathogens. Research in the
lab takes a multiscale approach ranging from experimental studies on the basic
biology of organisms to field studies across forest landscapes. Active collabo-
rations include projects with landscape ecologists, epidemiologists, molecular
biologists, entomologists, and forest managers. The primary research effort in
the lab is currently Phytophthora species in California coastal forests, with an
emphasis on sudden oak death. In conifer forests of the Sierra Nevada Mountains,
the lab studies a variety of diseases and their relationship to past and present
forest management and conservation issues. In addition to research, Dr. Rizzo
teaches undergraduate and graduate courses in mycology. Since 2004, he has
been director of the Science and Society program in the College of Agricultural
and Environmental Sciences. Dr. Rizzo also serves as the scientific advisor for
the California Oak Mortality Task Force.
Jim Stack, Ph.D., is director of the Great Plains Diagnostic Network (GPDN)
and a professor of plant pathology at Kansas State University. As director of
GPDN, Dr. Stack coordinates a nine-state project for the rapid detection and ac-
curate diagnosis of high-consequence pathogens and pests. He is the Principal
Investigator of a plant biosecurity project at the National Agriculture Biosecurity
Center and has collaborated on several international projects regarding plant
biosecurity. Prior to joining Kansas State, Dr. Stack was on the faculty at the
University of Nebraska and at Texas A&M University. He formerly worked for
EcoScience Corporation as the director of applied research, leading the discov-
ery, development, and commercialization of microbe-based products to protect
fruit from storage decay pathogens. His research interests include pathogen detec-
tion and diagnostics, pathogen ecology, and epidemiology.
Ché Weldon, Ph.D., M.Sc., holds research interests that have always been cen-
tered on amphibians, which started in 1997 with host–parasite interactions (B.Sc.
Honors project) and was followed by an in-depth assessment of the sustainable
global use of African clawed frogs (M.Sc. project, 1998–1999). The following
year, as a field biologist with the Southern African Frog Atlas and Red Data Proj-
ect, Dr. Weldon developed a more focused interest in amphibian conservation. It
was during this time that the amphibian chytrid fungus was first detected and a
global surge started to investigate its role in amphibian declines. For his studies
on amphibian chytrid in South Africa, Dr. Weldon received the W.O. Neitz medal
for best Ph.D. thesis in parasitology. With then-supervisor L. H. du Preez, Dr.
Weldon established the African Amphibian Conservation Research Group. Dr.
Weldon specialized in amphibian diseases in Africa as a Postdoctoral Fellowship
at North-West University, and later became a zoology lecturer at the university
where he has continued his research on the amphibian chytrid and amphibian
conservation. From the 12 peer-reviewed articles that Dr. Weldon has authored
in the past 6 years, 6 have been cited a total of 130 times.