28.03.2013 Views

Inoculum 63(3) - Mycological Society of America

Inoculum 63(3) - Mycological Society of America

Inoculum 63(3) - Mycological Society of America

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Newsletter <strong>of</strong> the <strong>Mycological</strong> <strong>Society</strong> <strong>of</strong> <strong>America</strong><br />

— In This Issue —<br />

Articles<br />

At the grave <strong>of</strong> William Alphonso Murrill<br />

Yeast Diversity Studied from the Gut <strong>of</strong> Australian<br />

Passalid Beetles<br />

MSA Business<br />

Secretary’s Report<br />

<strong>Mycological</strong> News<br />

Erast Parmasto, 1928-2012<br />

Items Needed for MSA Auction<br />

MSA Student Section<br />

Invitation to a Roundtable discussion: a Roadmap<br />

for Fungal Conservation Research in North <strong>America</strong><br />

Dominican Amber with Unknown Inclusions<br />

(help needed)<br />

MSA 2012 Abstracts<br />

<strong>Mycological</strong> Bookshelf<br />

Books in Need <strong>of</strong> Reviewers!<br />

Review: Waxcap Mushrooms <strong>of</strong> Eastern North<br />

<strong>America</strong><br />

Review: Biology <strong>of</strong> Lichens – Symbiosis, Ecology,<br />

Environmental Monitoring, Systematics and Cyber<br />

Applications<br />

<strong>Mycological</strong> Classifieds<br />

Mold and fungus testing and identification services<br />

Biological control, biotechnology and regulatory<br />

services<br />

Mycology On-Line<br />

Calendar <strong>of</strong> Events<br />

Sustaining Members<br />

— Important Dates —<br />

June 30<br />

Late registration deadline<br />

MSA annual meeting<br />

July 15<br />

Deadline for submission to <strong>Inoculum</strong> <strong>63</strong>(4)<br />

July 15-19<br />

MSA Annual Meeting<br />

Yale University<br />

July 31<br />

Registration Deadline<br />

EMBO Conference on Comparative Genomics<br />

<strong>of</strong> Eukaryotic Microorganisms<br />

Editor — Donald O. Natvig<br />

Department <strong>of</strong> Biology<br />

University <strong>of</strong> New Mexico<br />

Albuquerque, NM 87131<br />

Telephone: (505) 277-5977<br />

Fax: (505) 277-0304<br />

Email: dnatvig@gmail.com<br />

Supplement to<br />

Mycologia<br />

Vol. <strong>63</strong>(3)<br />

June 2012<br />

At the grave <strong>of</strong> William Alphonso<br />

Murrill<br />

On May 2, 2012 Tom Bruns, Rytas Vilgalys, Matt<br />

Smith (most recent Murrill successor at Florida), and their<br />

associates made a pilgrimage to the William Alphonso Murrill<br />

grave site. Shown in the photograph resulting from this<br />

trip are front row: Sunny Liao, Matt Smith, Zaiwei<br />

Ge, Jenny Talbot; back row: Laura Hubbard, Sara Branco,<br />

Rytas Vilgalys, Dylan Smith, Tom Bruns. An earlier photograph<br />

(also shown here) was made at the same site <strong>of</strong> Josiah<br />

Lowe, Robert L. Gilbertson, and Edson Setliff during the<br />

MSA – AIBS meeting, Gainesville, Florida, August 1985.<br />

William Alphonso Murrill (October 13, 1869-December 25,<br />

1957), one <strong>of</strong> the interesting personalities <strong>of</strong> mycology, was<br />

memorialized simply as he wished, as “author, naturalist,<br />

and editor” (which included serving as editor <strong>of</strong> Mycologia)<br />

(http://sciweb.nybg.org/science2/hcol/intern/murrill1.a<br />

sp). James W. Kimbrough, a successor to Murrill as a University<br />

<strong>of</strong> Florida mycologist, has researched Murrill’s life<br />

and published on him in the Mushroom Journal<br />

(http://www.mushroomthejournal.com/best<strong>of</strong>/articles.html)<br />

. The cemetery is the type locality for Xenosperma murrillii<br />

(Gilbertson, R. L., and M. Blackwell. 1987. Notes on<br />

wood-rotting fungi on junipers in the Gulf Coast Region. II.<br />

Mycotaxon 28:369-402.)<br />

Memorial information on Robert L. Gilbertson and a slide<br />

show including the photograph at the Murrill grave site can be<br />

found online at (http://lsb380.plbio.lsu.edu/Gil%20memorial.html).<br />

—Meredith Blackwell<br />

Louisiana State University<br />

Continued on following page


Yeast Diversity Studied from the Gut <strong>of</strong> Australian Passalid Beetles<br />

Hector Urbina, LSU graduate student from Caracas,<br />

was awarded an MSA Graduate Fellowship in<br />

2011, and he is very grateful to the <strong>Society</strong> for the<br />

award. MSA funding and a small grant from a program<br />

sponsored by the National Science Foundation<br />

and the Louisiana Board <strong>of</strong> Regents provided resources<br />

for a month in tropical forests in Queensland,<br />

Australia. The objective<br />

<strong>of</strong> the trip was to collect<br />

Australian passalid beetles<br />

(Passalidae,<br />

Coleoptera) and to isolate<br />

and purify yeasts<br />

from the gut <strong>of</strong> the<br />

wood-ingesting beetles.<br />

Strains associated with<br />

the gut <strong>of</strong> individuals in<br />

this family <strong>of</strong> the lignicolous<br />

insects are able to<br />

degrade and ferment a<br />

number <strong>of</strong> plant cell<br />

wall components.<br />

This family <strong>of</strong> insects<br />

mostly feed on decaying<br />

wood, and more<br />

than 30 species <strong>of</strong> passalids<br />

have been report-<br />

2 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

ed in from Australia. Yeasts from the Australian beetles<br />

as well as from the southeastern US, Guatemala,<br />

Panama, and Thailand will give Hector a snapshot <strong>of</strong><br />

the diversity <strong>of</strong> passalid-associated yeasts in several<br />

tropical regions <strong>of</strong> the world. Hector was successful<br />

in isolating more than 1000 yeasts from 200 adult<br />

Continued on following page<br />

Fig. 1. View <strong>of</strong> the Lamington National Park, Brisbane (Australia)


eetles in at least seven species <strong>of</strong> passalids. The beetle collecting<br />

was done under the guidance <strong>of</strong> his Australian colleagues,<br />

Roger Shivas, Justin Bartlett, Alistar McTaggant, and<br />

Desley Tree <strong>of</strong> the Department <strong>of</strong> Employment, Economic Development<br />

and Innovation, Queensland, Brisbane, Australia;<br />

and Dr. Owen Seeman, expert in Australian passalids and Collection<br />

Manager for Arachnida and Myriapoda, Biodiversity<br />

Program at the Queensland Museum. The research team visited<br />

a number <strong>of</strong> national parks including, Barron Gorge,<br />

Wooroonooran, Daintree, D’Agilar, Lamington, Main Rage and<br />

others. The beetle vouchers have arrived in Baton Rouge, and<br />

the yeasts are on their way to characterization. This work will<br />

increase our understanding <strong>of</strong> the diversity, phylogeny, and biochemical<br />

abilities <strong>of</strong> yeasts associated with the passalid gut.<br />

Hector also thanks his other Australian collaborators, Tomas<br />

Marney, who has helped in preparation <strong>of</strong> media and culturing<br />

<strong>of</strong> yeasts, and Dean Beasley for his help in depositing the yeast<br />

strains into the BRIP culture collection in Brisbane.<br />

Secretary’s Report<br />

Hello everyone! By the time you read this, summer<br />

will be upon us and we will be getting ready<br />

for the 2012 Annual Meeting at Yale University<br />

in New Haven, CT. Jean Lodge and the Program<br />

Committee have put together an exciting program<br />

for us this year; it will be a meeting to remember,<br />

and my final meeting as MSA Secretary. It has<br />

been an honor to serve!<br />

Council Business: Since my last report, the<br />

Council approved 6 actions by email poll. These<br />

included: Council support for Honorary Members<br />

and MSA Fellows selected by the Honorary<br />

Awards Committee; approval <strong>of</strong> Louise Glass for<br />

the Editorial Advisory Committee; and the approval<br />

<strong>of</strong> Tim James and Ignazio Carbone as Associate<br />

Editors <strong>of</strong> Mycologia.<br />

New Members: It is my pleasure to extend a<br />

warm welcome to new or returning members.<br />

Their membership will be formally approved at the 2012 Annual<br />

Business Meeting at Yale University.<br />

Canada - Tanay Bose, Michelle Hubbard, Hai Dt Nguyen, Judith<br />

Gagnon<br />

Czech Republic - Petr Baldrian<br />

Guyana - Dillon Husbands<br />

India- Krishnendu Acharya<br />

Mexico - Fidel Jaime<br />

—Hector Urbina<br />

MSA BUSINESS<br />

Jessie Glaeser,<br />

Secretary<br />

(Photo by Tom Volk)<br />

Fig. 2. Hector Urbina with a passalid collection<br />

at Jumrum Creek Conservation Park,<br />

Kuranda, Photo by Berndt Reinhard<br />

Netherlands - Maria Dam<br />

Sri Lanka - Dimuthu Manamgoda<br />

Thailand - Saranyaphat Boonmee<br />

Turkey - Mustafa Isiloglu<br />

United States - Suzette Senerez Arcibal, Silvia<br />

Bibbo, Kathryn E. Bushley, Lauren S C<strong>of</strong>fua,<br />

Sharifa Crandall, Andrea Davis, Wade Elmer,<br />

Karen Fisher, Serita D Frey, Erica Goldberger,<br />

Christine V Hawkes, Benjamin Held, Elizabeth<br />

Heppenheimer, Jaqueline X Hess, Miriam I<br />

Hutchinson, Eric Johnson, Justin Kaffenberger,<br />

Prasanna Kandel, Chinyere Knight , Melina<br />

Kozanitas, Samantha Lee, James Lendeme,<br />

Lotus A L<strong>of</strong>gren, Sara R Lopez, Shiloh<br />

Lueschow, Jing Luo, Benjamin Morgan, Eric<br />

Morrison, Kyle Patton, Eimy Plata, Elizabeth<br />

Roberts, Sarah Robinson, Katy Ryan, Jesse J Sadowsky,<br />

Jeffrey Shaw, Xiaomei Shu, Jason Slot,<br />

Daniel J Spakowicz, Iman Sylvain, Alexander Tice, Jeffrey<br />

Townsend, Gilberto Uribe Valdez, Roo W. Vandegrift, Gloria<br />

Wada, Eric Walberg, Holly Williams, Yuka Yajima, Kolea C<br />

Zimmerman<br />

Viet Nam - Hoang Pham<br />

Emeritus Members: There were no applications for Emeritus<br />

status in the past two months. Emeritus status is granted to<br />

MSA members who are retired from active pr<strong>of</strong>essional em-<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 3


ployment and who have been MSA members for 15 years or<br />

more. There is no membership fee for Emeritus members although<br />

there is a reduced fee for access to Mycologia. If you<br />

are interested in becoming an Emeritus member, please contact<br />

me directly at msasec01@yahoo.com.<br />

Deceased Members: We are saddened to hear <strong>of</strong> the passing<br />

<strong>of</strong> Erast Parmasto, an Honorary Member <strong>of</strong> MSA. Historian<br />

Ron Petersen has information about Dr. Parmasto’s career later<br />

in this issue. We were also sorry to learn, belatedly, <strong>of</strong> the passing<br />

<strong>of</strong> B.T. Lingappa, Dexter H. Howard, and Marjorie Anchel.<br />

Reminder: Renewing your MSA membership is easier than<br />

ever! Just log in to the MSA website at http://www.msafungi.org.<br />

There is now an email reminder system available if you<br />

have forgotten your MSA user id or password.<br />

REMINDER: MSA Directory Update: Is your information<br />

up-to-date in the MSA directory? The <strong>Society</strong> is relying more<br />

and more on email to bring you the latest MSA news, awards<br />

announcements and other timely information. To ensure that<br />

4 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

you receive <strong>Society</strong> blast emails and the <strong>Inoculum</strong> as soon as it<br />

comes out, and so that your colleagues can keep in touch,<br />

please check the accuracy <strong>of</strong> your email address and contact information<br />

in the online directory. This can be accessed via our<br />

web site at www.msafungi.org. If you need assistance with<br />

updating your membership information, please contact our Association<br />

Manager at Allen Press, the always-helpful Kay Rose<br />

at krose@allenpress.com.<br />

Please do not hesitate to contact me about MSA Business or any<br />

questions that you may have about the <strong>Society</strong>. In recent years<br />

we have suffered an alarming decline in membership and it<br />

would be wonderful to reverse this trend. The first step is for<br />

everyone who is currently a member to renew for the upcoming<br />

year. And don’t forget to recommend MSA to your amateur or<br />

pr<strong>of</strong>essional colleagues who are interested in fungi – be they<br />

pathologists, geneticists, ecologists, or people who just like to<br />

wander around in the woods. There is room in MSA for all!<br />

MYCOLOGICAL NEWS<br />

Erast Parmasto, 1928-2012<br />

The MSA has lost one <strong>of</strong> its honorary members, Erast<br />

Parmasto <strong>of</strong> Tallin, Estonia.<br />

Erast was an academician <strong>of</strong> the Academy <strong>of</strong> Sciences<br />

<strong>of</strong> Estonia and expert in biosystematics, especially in the taxonomy<br />

<strong>of</strong> Hymenomycetes and statistics <strong>of</strong> spore production<br />

and measurement. An outstanding student in Estonian<br />

schools during World War II, Erast thereafter became a student<br />

<strong>of</strong> Apollinari Bondarzew in Leningrad (now St. Petersburg),<br />

thereby establishing a promising foundation on which<br />

to build a career, were it not the early days <strong>of</strong> the “Cold War”<br />

and difficult communication outside the Soviet hegemony.<br />

By 1972, Erast was elected a member <strong>of</strong> the Estonian Academy<br />

<strong>of</strong> Sciences, at the beginning <strong>of</strong> a career spent at the Institute<br />

<strong>of</strong> Biology <strong>of</strong> the Estonian Academy <strong>of</strong> Sciences in<br />

Tartu, Estonia.<br />

Like many <strong>of</strong> his age and geographic location, Erast<br />

lived through the days <strong>of</strong> Germany occupation, followed by<br />

years <strong>of</strong> Soviet Russian domination. During the latter<br />

period, it was almost impossible to gain recognition be-<br />

yond the “Iron Curtain,” but Erast persisted in publishing<br />

(always correctly in Russian) scientific papers, including<br />

several major monographs and, when possible,<br />

putting them in the hands <strong>of</strong> workers outside the Russian<br />

hegemony.<br />

A personal note: I first met Erast in Stockholm,<br />

Sweden. He had been invited to join a foray to the<br />

Swedish north, and we spent hours on the train from<br />

Stockholm to Umeå. Although we plied him with questions<br />

about conditions in Estonia, he would not answer<br />

directly, implying that “the walls have ears.” Even on<br />

the ferry he would not comment on such things. At last<br />

we reached our destination near the Norwegian border<br />

—Jessie A. Glaeser<br />

MSA Secretary<br />

msasec1@yahoo.com<br />

and the following morning we climbed a nearby hill and rested<br />

on a large rock, where Erast, as was his habit, lit a cigarette.<br />

There, surrounded for many meters by nothing by nature,<br />

he <strong>of</strong>fered, “Now I will tell you about conditions in<br />

Estonia.” And they were grim but revelatory to those <strong>of</strong> us<br />

working in freer, less ideological circumstances. Only relatively<br />

recently did life become appreciably more comfortable<br />

for science in the Baltic States.<br />

Author <strong>of</strong> over 400 scientific papers, an obituary with<br />

photos is available at http://erast.ut.ee/ErastParmasto/. He is<br />

survived by his wife and scientific collaborator, Ilmi. Erast<br />

will be remembered affectionately for his dry wit (necessary<br />

for survival in his lifetime) and careful attention to scientific<br />

detail. He mentored numerous younger colleagues who will<br />

carry on with the benefit <strong>of</strong> his experience.<br />

MSA Auction Items<br />

—Ron Petersen<br />

University <strong>of</strong> Tennessee<br />

The MSA auction (Wednesday evening, July 18) is one <strong>of</strong><br />

the major fundraising efforts <strong>of</strong> the society. All donations to the<br />

auction, silly or serious, are very welcome. If you plan to donate<br />

an item (or items) to the auction, please bring them with you<br />

and drop me an e-mail to let me know what you are bringing. If<br />

you don’t plan to come this year but still would like to donate,<br />

please send items to the address below.<br />

—Karen Hughes<br />

Ecology and Evolutionary Biology<br />

Hesler Biology Building Rm 330<br />

University <strong>of</strong> Tennessee, Knoxville 37920


MSA Student Section<br />

At the 2012 meeting this year in<br />

New Haven, the <strong>Mycological</strong> <strong>Society</strong><br />

<strong>of</strong> <strong>America</strong> will be launching a Student<br />

Section. This student-run group<br />

within the MSA will provide opportunities<br />

for students to network with<br />

other students in their own fields and<br />

beyond. It will also be valuable for<br />

student members <strong>of</strong> MSA seeking<br />

connections with those performing<br />

cutting edge research in mycology<br />

and thus has the potential to inspire<br />

future collaborative research. The<br />

Student Section is open and inclusive;<br />

we welcome the participation <strong>of</strong> all<br />

students (and faculty!) in building this<br />

group. We look forward to your participation<br />

in our new mentorship program<br />

and hope you can also join us at<br />

the following events.<br />

Sunday July 15 MSA Student Workshop:<br />

Using fungi to get a job<br />

Join us as Harvard University mycology pr<strong>of</strong>essor Anne<br />

Pringle leads our first MSA Student Section event. This pr<strong>of</strong>essional<br />

development workshop will take place after the<br />

Foray and before the Welcome Reception on Sunday.<br />

Abstract. Some ads specifically ask for mycologists,<br />

more <strong>of</strong>ten, departments are looking for conceptually oriented<br />

biologists. Fungi <strong>of</strong>fer myriad opportunities for connecting<br />

to broader themes in ecology, evolution, and other fields;<br />

making those connections can be critical to a successful job<br />

hunt. I’ll talk about my own experiences on the job market<br />

and share the research and teaching statements I used to find<br />

my current position. We’ll then practice job strategies by tak-<br />

Invitation to a Roundtable Discussion<br />

Toward a Roadmap for Fungal Conservation<br />

Research in North <strong>America</strong>: Tools,<br />

Data Needs and Collaborations<br />

MSA Annual Meeting<br />

Monday 16 July, 4-6 PM<br />

The 80 th annual meeting <strong>of</strong> the MSA will include a<br />

roundtable discussion on the tools, data, and collaborations<br />

that are needed for fungal conservation in North <strong>America</strong>.<br />

The threat <strong>of</strong> large-scale species extinctions during this<br />

century makes understanding the extent, functions, and consequences<br />

<strong>of</strong> loss <strong>of</strong> biodiversity one <strong>of</strong> the most imperative<br />

areas <strong>of</strong> biological research <strong>of</strong> our time. At the same time<br />

that environmental DNA studies highlight the depths <strong>of</strong> unknown<br />

fungal diversity, taxonomic specialists are continuing<br />

to decrease in number. And while studies <strong>of</strong> fungal community<br />

ecology are abundant, there have been very few studies<br />

ing the time to talk about our research<br />

“mission statements” and<br />

“cocktail-party tidbits”. I’ll explain<br />

why these are useful and we’ll finish<br />

by sharing our newly crafted statements<br />

with each other.<br />

Monday July 16th: First<br />

Annual Student Section<br />

Mixer<br />

Join us before the poster session<br />

as we gather in the poster area for a<br />

casual meet and greet with c<strong>of</strong>fee,<br />

dessert, and other drinks provided.<br />

The mixer will go from 6:00-6:30<br />

p.m.<br />

Mentorship Program<br />

Mentorship programs are an excellent<br />

way to introduce students to<br />

other researchers with similar interests.<br />

In addition, mentors may assist in steering student research<br />

or by providing information about previous studies or<br />

future opportunities.<br />

For these reasons, all students who expressed interest on<br />

their registration form have been assigned a senior mycologist<br />

who will serve as their mentor for the duration <strong>of</strong> the<br />

New Haven meeting. Students will meet their senior mycologist<br />

mentor at the Student Section Mixer on Monday, July<br />

16th.<br />

We appreciate your involvement in the commencement<br />

<strong>of</strong> the new MSA Student Section and look forward to seeing<br />

you all in New Haven!<br />

—Mia Maltz<br />

Department <strong>of</strong> Ecology and Evolutionary Biology<br />

University <strong>of</strong> California, Irvine<br />

in North <strong>America</strong> conducted with an explicit conservation<br />

biology focus; the large-scale Survey and Manage program<br />

<strong>of</strong> the U.S. Forest Service under the Northwest Forest Plan<br />

(Molina 2008) is a rare example. The extent <strong>of</strong> species endemism<br />

and rarity is unknown for most habitat types, and<br />

fungi have not been considered for conservation action (e.g.,<br />

RED data lists) in the U.S.A.<br />

Our goal for this roundtable is to consider ways to make<br />

progress in fungal conservation by considering the tools, data<br />

needs, and collaborations necessary to place fungi in a conservation<br />

framework comparable to that which exists for plant<br />

and animal species. In terms <strong>of</strong> tools, potential topics include<br />

harnessing molecular tools, training new taxonomic specialists,<br />

and informatics tools for information gathering and transfer<br />

for herbarium collections and ecological and distribution<br />

data. Potential types <strong>of</strong> data necessary for conservation in-<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 5


clude threat assessments; taxonomic identifications; distributions;<br />

life cycle and fruiting patterns; habitat requirements;<br />

species interactions; diversity, endemism and the characteristics<br />

<strong>of</strong> fungal biodiversity hotspots; and the dynamics inherent<br />

to many <strong>of</strong> these data. We have to consider how to integrate<br />

existing conservation-relevant, but not explicitly<br />

conservation-oriented, data into a clearer conservation framework;<br />

for example, whether inferences about fungal distribution<br />

and occurrence can be made from studies <strong>of</strong> ecological<br />

community composition based on small plots (e.g., recent<br />

studies using high-throughput environmental DNA sequencing).<br />

Potential types <strong>of</strong> collaboration include stronger networks<br />

between ecologists, taxonomists, and population geneticists,<br />

but also pr<strong>of</strong>essional-amateur collaborations as a<br />

means to address the “dwindling taxonomists” issue.<br />

An anticipated result <strong>of</strong> the roundtable will be increased<br />

clarity on whether to – and if so, how to – bring greater attention<br />

to fungal conservation and lay out a roadmap for a<br />

collaborative, scientifically rigorous way forward – specific<br />

areas in which we need to concentrate our research efforts,<br />

how to set priorities, how to integrate knowledge from many<br />

different sources, how to include conservation issues in our<br />

Dominican Amber with Unknown Inclusions<br />

Can you identify the inclusion? Trapped in a Dominican<br />

amber specimen is a cluster <strong>of</strong> what looks<br />

like fungi, including one that is suspended in<br />

the amber. There appears to be five gilled clusters,<br />

which are partially exposed on the back <strong>of</strong> the piece.<br />

I would like assistance in classifying these inclusions.<br />

Please contact me at mail.acton@gmail.com if<br />

you can <strong>of</strong>fer guidance in identifying these mystery<br />

inclusions.<br />

6 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

—Jeanette Acton<br />

other research (Molina et al. 2011), and how to educate, inform<br />

and involve the general public.<br />

We invite all attenders <strong>of</strong> the upcoming annual meeting<br />

to attend the roundtable discussion, ready to bring your<br />

thoughts and experiences to these issues. Invited facilitators<br />

will provide brief (approximately 5-minute) introductions on<br />

several topics (see the meeting website,<br />

http://msa2012.net/schedule/symposia.php#roadmap, for<br />

more details) to stimulate discussion, but for the main part it<br />

will be an ‘open mike’ format with active participation <strong>of</strong><br />

anyone in attendance. We hope to see you there.<br />

—On behalf <strong>of</strong> the MSA Conservation Committee:<br />

Todd Osmundson and Else Vellinga<br />

University <strong>of</strong> California, Berkeley<br />

REFERENCES<br />

Molina R. 2008. Protecting rare, little known, old-growth forest-associated<br />

fungi in the Pacific Northwest USA: a case<br />

study in fungal conservation. <strong>Mycological</strong> Research 112:<br />

613-<strong>63</strong>8.<br />

Molina R, Horton TR, Trappe JM, Marcot BG. 2011. Addressing<br />

uncertainty: How to conserve and manage rare or<br />

little-known fungi. Fungal Ecology 4: 134-146.


MSA 2012 ABSTRACTS<br />

Adenipekun, Clementina O 1 , Omasan E Ejoh 2 , and Adeniyi A Ogunjobi 2 .<br />

1 Department <strong>of</strong> Botany, University <strong>of</strong> Ibadan, Ibadan, Nigeria, 2 Department <strong>of</strong><br />

Microbiology, University <strong>of</strong> Ibadan, Ibadan, Nigeria. Effect <strong>of</strong> Pleurotus tuberregium<br />

Singer and microorganisms on degradation <strong>of</strong> unsterilized soil contaminated<br />

with cutting fluids<br />

The study <strong>of</strong> Pleurotus tuber-regium Singer and indigenous microorganisms<br />

isolated from unsterilized soil polluted with cutting fluids were investigated<br />

over an incubation period <strong>of</strong> 0 and 2 months. The ability <strong>of</strong> these organisms to degrade<br />

Total Petroleum hydrocarbon (TPH) present in the cutting fluids, lignin<br />

content <strong>of</strong> rice straw, their enzyme activity as well as their ability to accumulate<br />

heavy metals present in the polluted soils were monitored. In all the soil samples<br />

(unsterilized soil polluted with cutting fluids (US) and unsterilized soil polluted<br />

with cutting fluids incubated with P. tuber-regium (USP), a significant increase in<br />

the nutrient content was observed with USP recording the highest. The heavy<br />

metal content <strong>of</strong> the three soil samples decreased with increase in incubation period<br />

showing that bioaccumulation <strong>of</strong> the heavy metals had occurred. The indigenous<br />

microbes alone in US accumulated heavy metals better than the indigenous<br />

microbes in USP. TPH loss (36.23%) was recorded at 10% cutting fluids concentration<br />

in USP compared to US with 19.6%. The unsterilized contaminated<br />

soil inoculated with P. tuber-regium recorded the highest lignin degradation as the<br />

polyphenol oxidase and peroxidase activity <strong>of</strong> the organisms in all the samples<br />

showed a gradual increase. The microorganisms isolated from unsterilized soil<br />

samples included Bacillus licheniformis, B. cereus, Bacillus sp., Pseudomonas<br />

aeruginosa, Pseudomonas sp., Corynebacterium sp., Neisseria sp., Trichoderma<br />

harzanium, Aspergillus niger, A. flavus, A. glaucus, Mucor sp. and Rhizopus sp.<br />

Acharya, Krishnendu 1 , Arun K Dutta 1 , Prakash Pradhan 1 , and Anirban Roy 2 1<br />

.<br />

Molecular and Applied Mycology and Plant Pathology Laboratory, Department<br />

<strong>of</strong> Botany, University <strong>of</strong> Calcutta, Kolkata, INDIA, 2 West Bengal Biodiversity<br />

Board, Paribesh Bhawan, Kolkata, INDIA. Habitat diversity <strong>of</strong> macr<strong>of</strong>ungi in<br />

the Indian part <strong>of</strong> Sundarban<br />

The Sundarban Biosphere Reserve covers an area <strong>of</strong> approximate 25,500<br />

sq km. However, only 9,<strong>63</strong>0 sq km is in India, with the rest located in Bangladesh.<br />

This area in India is demarcated by the Hoogly River in the west, the Bay-<strong>of</strong>-Bengal<br />

in the south, and the Harinbhanga and Raimangal Rivers in the east and the<br />

Dampier-Hodges line in the north. The wide range <strong>of</strong> phyto-topographical features,<br />

presence <strong>of</strong> spatial and temporal variability in hydrological regimes and the<br />

diverse substrata has contributed to the high biodiversity <strong>of</strong> this region. Mangrove<br />

forest dominates in the core area and some margins <strong>of</strong> the deltic island, where inhabited<br />

lands show typical coastal vegetation. Several anthropogenic activities,<br />

recurring natural calamities and recent incidences <strong>of</strong> cyclones (e.g., Aila) have<br />

added to the vulnerability <strong>of</strong> this biologically rich world heritage site. The climate<br />

is characterized by relatively high temperature and humidity (>80%) throughout<br />

the year, and well-distributed rainfall during the monsoon season. Temperatures<br />

rise from a daily minima <strong>of</strong> 10ºC in the winter to a maximum <strong>of</strong> about 43ºC in<br />

March and may exceed 32ºC during the monsoon season. The average annual<br />

rainfall in the Indian part <strong>of</strong> the Sundarban region is 1,662 mm. During our survey<br />

(2010-2011), 77 quadrates were studied each measuring 20 _ 20 meters.<br />

Macr<strong>of</strong>ungi were collected, photographed, identified and were preserved with accession<br />

numbers. A total <strong>of</strong> 71 species <strong>of</strong> macr<strong>of</strong>ungi belonging to 29 families, 53<br />

genera were identified and their ecology was evaluated.<br />

Afshan, Najam-ul-Sahar, Abdul Nasir Khalid, and A R Niazi. Centre for Undergraduate<br />

Studies, University <strong>of</strong> the Punjab, Quaid-e-Azam Campus, Lahore,<br />

54950, Pakistan. Rust fungi <strong>of</strong> Himalayan Moist Temperate forests <strong>of</strong> Pakistan,<br />

their diversity and distribution: An overview<br />

Himalayan Moist Temperate forests <strong>of</strong> Pakistan have coniferous species,<br />

mixed evergreen and deciduous patches and broad leaved forests. These forests <strong>of</strong><br />

Pakistan merge downward with the tropical thorn forests and upwards with the<br />

alpine meadows. The precipitation may exceed 600 mm, thus making <strong>of</strong> these<br />

montane forests a bio-geographical crossroad between submontane and alpine<br />

meadow vegetation. Being rich in plant diversity, these forests harbor a large<br />

number <strong>of</strong> rust fungi that are obligate parasites <strong>of</strong> plants. Although the fungal flora<br />

<strong>of</strong> Pakistan has been explored by several workers in the past, the important group<br />

<strong>of</strong> rust fungi has largely been neglected resulting in a paucity <strong>of</strong> literature and very<br />

fragmentary knowledge <strong>of</strong> these fungi, particularly in Himalayan Moist Temperate<br />

forests <strong>of</strong> Pakistan. This study was undertaken to explore and assess the diversity<br />

and distribution <strong>of</strong> rust fungi along with their respective host plants in this<br />

floristically rich area. This preliminary study <strong>of</strong> the hosts and their rusts includes<br />

approximately 169 species and 19 genera <strong>of</strong> rust fungi on approximately 260<br />

species <strong>of</strong> host plants. These rust fungi include one species each <strong>of</strong> genus Cronartium,<br />

Hyalopsora, Miyagia, Monosporidium, Pucciniastrum, Pucciniostele,<br />

Thekopsora and Uredinopsis respectively; two species each <strong>of</strong> Gymnosporangium,<br />

Peridermium, Pucciniastrum and Uredo; three and four species <strong>of</strong><br />

Caeoma and Coleosporium respectively; seven <strong>of</strong> Melampsora; nine <strong>of</strong> Aecidium;<br />

eleven <strong>of</strong> Phragmidium; twenty four species <strong>of</strong> Uromyces and ninety five<br />

species <strong>of</strong> largest genus <strong>of</strong> rust fungi, Puccinia. This study enlists rust fungi <strong>of</strong> Himalayan<br />

Moist Temperate forests <strong>of</strong> Pakistan and presents their species richness<br />

and geographical distribution. This work will help to prepare a checklist <strong>of</strong> rust<br />

fungi <strong>of</strong> this area and will ultimately lead to the documentation and preparation <strong>of</strong><br />

monograph <strong>of</strong> the rust fungi <strong>of</strong> Pakistan.<br />

Ahrendt, Steven R and Jason E Stajich. Department <strong>of</strong> Plant Pathology and Microbiology,<br />

University <strong>of</strong> California, Riverside, CA 92521. Analysis <strong>of</strong> a putative<br />

sensory rhodopsin in the Chytridiomycota<br />

Rhodopsin is a seven-transmembrane G-protein coupled receptor<br />

(GPCR) that responds to light through photoisomerization <strong>of</strong> a covalently bound<br />

11-cis-retinal molecule. This action forces helical motion and subsequent signal<br />

transduction through activation <strong>of</strong> the coupled G-protein. Rhodopsin homologs<br />

have only been identified in metazoan lineages, and while Dikarya fungi are<br />

known to have opsins, phytochromes, and cryptochromes to sense light, this type<br />

<strong>of</strong> rhodopsin has only been found encoded in the genomes <strong>of</strong> the early diverging<br />

Chytridiomycota and Blastocladiomycota fungi. Previous and current work has<br />

shown that zoosporic fungi are phototaxic but the molecular mechanisms <strong>of</strong> light<br />

sensing in early diverging fungi has not been explored. Here we describe structural<br />

and functional analyses <strong>of</strong> rhodopsin proteins identified in two species <strong>of</strong><br />

Chytridiomycota, the amphibian pathogen Batracochytrium dendrobatidis and<br />

the terrestrial saprotroph Spizellomyces punctatus. Comparative genomics analyses<br />

<strong>of</strong> rhodopsin and flagellum genes in the Chytridiomycota, Zygomycetes, and<br />

Dikarya show a correlation <strong>of</strong> flagella and rhodopsin presence across the fungi.<br />

Computational modeling <strong>of</strong> the B. dendrobatidis and S. punctatus proteins indicates<br />

that they both adopt the seven transmembrane helix formation typical <strong>of</strong><br />

rhodopsin proteins. Additionally, structural features associated with rhodopsin<br />

proteins are observed as well. The B. dendrobatidis protein sequence is notably<br />

lacking the conserved lysine residue, however this residue is present in the S.<br />

punctatus sequence. Phototaxis assays were performed to determine in vivo function<br />

<strong>of</strong> the rhodopsin proteins. Ongoing work to characterize protein function<br />

through expression in Pichia pastoris may provide insight into the mechanism <strong>of</strong><br />

retinal binding and photoisomerization in these chytrid rhodopsins.<br />

Aime, M Catherine. Louisiana State University Agricultural Center, Department<br />

<strong>of</strong> Plant Pathology and Crop Physiology, Baton Rouge, LA 70803. Basidiomycete<br />

taxonomy without dual nomenclature<br />

The new International Code <strong>of</strong> Nomenclature for algae, fungi, and plants<br />

(Melbourne Code) discontinued the application <strong>of</strong> dual nomenclature in Fungi.<br />

Under earlier Codes Article 59 permitted more than one name for pleomorphic<br />

non-lichenized basidiomycetes. Effective immediately all names will compete for<br />

priority. As <strong>of</strong> January 1, 2013, publication <strong>of</strong> alternative names for species or<br />

genera will not be allowed. These changes affect basidiomycete taxonomy in several<br />

ways. To make one example, anamorphic yeast states occur in all three subphyla<br />

<strong>of</strong> Basidiomycota. Under previous versions <strong>of</strong> the Code these were treated<br />

separately from teleomorphic species and assigned to form genera principally<br />

based on carbon assimilation tests. Molecular data are now used to place these<br />

within a phylogenetic framework, highlighting the artificiality <strong>of</strong> the old system.<br />

For example, species <strong>of</strong> Sporobolomyces occur across most <strong>of</strong> the yeast-forming<br />

Pucciniomycotina classes and species <strong>of</strong> Rhodotorula can be found across both<br />

Ustilaginomycotina and Pucciniomycotina, although the type species for both<br />

genera are placed in Sporidiobolales. Thus restriction <strong>of</strong> generic names to include<br />

only those species within a monophyletic lineage will dictate the necessity <strong>of</strong><br />

many name changes within these groups as extra-Sporobolomyces, -Rhodotorula<br />

and other anamorphic basidiomycete yeast species are integrated into a phylogenetic-based<br />

taxonomy. These and many other taxonomic challenges for the community<br />

will be discussed as they apply to different groups <strong>of</strong> impacted Basidiomycota.<br />

Ainsworth, A Martyn 1 , David Parfitt 2 , Hilary J Rogers 2 , and Lynne Boddy 2 1<br />

.<br />

Royal Botanic Gardens, Kew / Natural England, Mycology Section, Jodrell<br />

Lab., Kew. TW9 3AB. United Kingdom, 2 Cardiff University, Cardiff School <strong>of</strong><br />

Biosciences, Cardiff. CF10 3AX. United Kingdom. Species concepts in European<br />

stipitate hydnoid fungi<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 7


The standard taxonomic treatment <strong>of</strong> European stipitate hydnoid species<br />

(ectomycorrhizal tooth fungi in the genera Bankera, Hydnellum, Phellodon and<br />

Sarcodon), written by R.A. Maas Geesteranus, is almost 40 years old. It is timely<br />

therefore to re-examine the morphological species concepts therein using the<br />

results <strong>of</strong> a combined molecular and morphological approach. These fungi have<br />

received conservation-related publicity across Europe and are now generally regarded<br />

as nitrogen-sensitive tree symbionts frequently associating with roots <strong>of</strong><br />

Fagaceae and Pinaceae. A decline in European fruiting populations is mainly ascribed<br />

to increased habitat loss and aerial nitrogen deposition. Of the 18 extant<br />

stipitate hydnoid species currently included on the British and Irish checklist, 15<br />

are recognised as <strong>of</strong> conservation importance (priority BAP species) in the UK.<br />

However all conservation status assessments currently depend on morphological<br />

taxonomic concepts and accurate identification <strong>of</strong> fruit bodies. Recognising that a<br />

mushroom is a stipitate hydnoid is <strong>of</strong>ten straightforward, but naming the species<br />

on existing fruit body criteria can be fraught with difficulty. This is due to a combination<br />

<strong>of</strong> identification and taxonomic issues. In order to improve assessments<br />

<strong>of</strong> species distributions, conservation status and ecological/conservation requirements,<br />

we have adopted a combined approach using molecular (ITS1 sequencing)<br />

and traditional (fruit body and spore morphology) methods. We have discovered<br />

that the list <strong>of</strong> extant taxa is rather longer (and increasing) than the list <strong>of</strong> currently<br />

accepted names. There are undoubtedly cryptic and not-so-cryptic hydnoid taxa<br />

to be described in Britain and elsewhere in Europe. We are now comparing<br />

species names with sequence-based groupings in order to pin-point the undescribed<br />

species. This combined approach will facilitate stipitate hydnoid identification,<br />

help to locate their “best” sites for conservation and accelerate their belowground<br />

ecological study.<br />

Albu, Sebastian, Tomas A Rush, and M Catherine Aime. Louisiana State University<br />

Agricultural Center, Department <strong>of</strong> Plant Pathology and Crop Physiology,<br />

Baton Rouge, Louisiana, 70803. Description <strong>of</strong> two anamorphic yeasts in the<br />

Ustilaginales<br />

Two basidiomycete yeasts belonging to the Ustilaginales were isolated in 2011<br />

from the leaves <strong>of</strong> several fern species in Baton Rouge, Louisiana. Based on a<br />

combination <strong>of</strong> assimilation tests and phenotypic characterization, Farysizyma sp.<br />

nov. SA209 (Anthracoideaceae) and Pseudozyma sp. nov. SA575 (Ustilaginaceae)<br />

represent previously undescribed anamorphs in the Ustilaginomycetes.<br />

Colony morphologies are initially yeast-like, subsequently developing pseudohyphal<br />

tufts around the growing margin. Maximum likelihood analyses using four<br />

loci, the internal transcribed spacer (ITS) region, large subunit (LSU) and small<br />

subunit <strong>of</strong> the nuclear rDNA cistron, and translation elongation factor 1-alpha indicate<br />

that both isolates belong within Ustilaginales. The LSU and ITS regions <strong>of</strong><br />

these isolates were compared to sequences <strong>of</strong> other available Farysizyma and<br />

Pseudozyma anamorphs and to related teleomorphs. Farysizyma sp. nov. SA209<br />

is part <strong>of</strong> a Farysizyma/Farysia clade in Anthracoideaceae containing all other<br />

known species <strong>of</strong> Farysizyma. Pseudozyma sp. nov. is sister to Sporisorium<br />

hwangense within a larger clade <strong>of</strong> predominantly Sporisorium species that includes<br />

the type (S. sorghi). Comparison <strong>of</strong> our data with phenotypic descriptions<br />

and sequence data from all known species <strong>of</strong> Farysizyma and Pseudozyma indicates<br />

that neither isolate has been previously described in the anamorphic yeast<br />

state. However, without sequence information from all known members <strong>of</strong> Ustilaginales,<br />

we cannot rule out the possibility that either <strong>of</strong> these isolates may represent<br />

a previously described teleomorphic Farysia or Ustilago/Sporisorium<br />

species.<br />

Allen, Michael F. Center for Conservation Biology, University <strong>of</strong> California,<br />

Riverside, CA 92521 USA. Mycorrhizae and resource acquisition: dynamics<br />

in fluctuating environments<br />

Mycorrhizae exist from organic layers deep into soil pr<strong>of</strong>iles, even into fracturing<br />

bedrock. As soils dry, deeper water either from groundwater or from water bound<br />

in cracks and pockets is acquired and utilized or reallocated via hydraulic redistribution.<br />

N is acquired from surface soils even under drought conditions. We utilized<br />

shifts in natural abundance isotope ratios to better understand spatial and<br />

temporal patterns <strong>of</strong> resource acquisition. Subsequently, utilizing continuous sensor<br />

and observation platforms, we monitored shifts in roots, mycorrhizal fungi,<br />

and respiration in response to changing temperature and moisture. Acute perturbations<br />

consisting <strong>of</strong> either large storms or drought result in plant mortality, but<br />

also subsequently re-order root and fungal symbiosis shifting both the types and<br />

spatial structure <strong>of</strong> resource acquisition. This may be because in seasonal environments,<br />

root and fungal activity show complex and <strong>of</strong>ten rapid responses to environmental<br />

change. We can utilize these shifts to develop a better understanding<br />

<strong>of</strong> belowground response dynamics to more chronic climate change or to acute<br />

perturbations such as hurricanes or drought.<br />

Annis, Seanna L 1 , Rafael Garcia 2 , Beth Calder 2 , and Kathryn L Hopkins 3 1 2<br />

.<br />

School <strong>of</strong> Biology and Ecology, University <strong>of</strong> Maine, Orono, ME, 04469, Department<br />

<strong>of</strong> Food Science and Human Nutrition, University <strong>of</strong> Maine, Orono,<br />

8 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

ME, 04469, 3 Cooperative Extension, University <strong>of</strong> Maine, Orono, ME, 04469.<br />

Identification <strong>of</strong> fungal contamination in bottled maple syrup<br />

Maple syrup processors in the northeastern USA occasionally observe fungal contamination<br />

in their bottled maple syrup containers. The concern is that the contaminants<br />

may pose a health risk. We identified fungal organisms in 32 bottles <strong>of</strong><br />

syrup from different processors in the northeastern USA. Most containers had one<br />

fungus, but some were contaminated with multiple fungi. Fungi were identified<br />

by morphology and DNA sequences <strong>of</strong> their ribosomal internal transcribed spacer<br />

regions and B tubulin genes. Multiple species <strong>of</strong> the genera Penicillium and Aspergillus,<br />

single species <strong>of</strong> Wallemia, and as yet unidentified yeasts were isolated<br />

from multiple bottles. Species <strong>of</strong> Paecolimyces and Cladosporium were isolated<br />

from single bottles <strong>of</strong> syrup. Some <strong>of</strong> the genera <strong>of</strong> Penicillium and Aspergillus<br />

identified are known to produce mycotoxins, and the production <strong>of</strong> these compounds<br />

in maple syrup is being evaluated. Syrup is typically bottled at 82 C to decrease<br />

the risk <strong>of</strong> microbial contamination. Spores from some <strong>of</strong> the fungi were<br />

able to germinate after treatment at 70 C for 3 minutes. A higher bottling temperature<br />

may be one <strong>of</strong> the changes in bottling practices required to minimize fungal<br />

contamination. This research will result in new recommendations to maple syrup<br />

processors on preventing future fungal contamination <strong>of</strong> bottled syrup.<br />

Baldrian, Petr, Tomás Vetrovsky, Jana Vorísková, Ivana Eichlerová, Jaroslav<br />

Snajdr, Lucia Zifcáková, and Martina Stursová. Laboratory <strong>of</strong> Environmental Microbiology,<br />

Institute <strong>of</strong> Microbiology <strong>of</strong> the ASCR, Prague, Czech Republic. Exploring<br />

fungal community structure and function in forest soils: challenges<br />

and limitations <strong>of</strong> current methodologies<br />

Presently, the structure and function <strong>of</strong> soil fungal communities receives considerable<br />

attention. This attention is fuelled by the recognition <strong>of</strong> the key role <strong>of</strong> fungi<br />

in the C and N cycling in soils, especially <strong>of</strong> the forest biomes. In addition, the recent<br />

establishment <strong>of</strong> high-throughput-sequencing methods, labelling with stable<br />

isotopes or metaproteomics <strong>of</strong>fers a much higher resolution <strong>of</strong> the current studies.<br />

The interpretation <strong>of</strong> experimental results is, however, still challenging due to the<br />

fact that many methods may potentially contain more or less apparent biasses.<br />

This contribution aims to point at the most important limitations <strong>of</strong> current<br />

methodologies to explore fungal abundance, community composition and function<br />

in forest soils as studied using biomass quantification techniques (PLFA, ergosterol,<br />

qPCR), shotgun or amplicon-based next-generation-sequencing, stable<br />

isotope probing and environmental metaproteomics. In the case <strong>of</strong> microbial biomass<br />

surveys, our results show that various methods (rDNA quantification, PLFA<br />

or ergosterol assays) yield widely different results <strong>of</strong> the fungal biomass content<br />

or fungal/bacterial biomass ratio and identify the differences in the composition<br />

<strong>of</strong> fungal mycelia / genomes as a source <strong>of</strong> such errors. For the next-generationsequencing<br />

data, we show that shotgun methods are not consistent with PCRbased<br />

methods and that the use <strong>of</strong> rDNA markers is highly biased due to uneven<br />

content <strong>of</strong> ITS copies per fungal genome as can be demonstrated if single-copy<br />

genes are sequenced. The approaches to explore the active fraction <strong>of</strong> the total<br />

fungal community by the analysis <strong>of</strong> RNA-derived sequencing or the use <strong>of</strong> Stable<br />

Isotope Probing and, most recently, environmental metaproteomics <strong>of</strong>fer an<br />

attractive insight into the functioning <strong>of</strong> fungal communities. However, even the<br />

use <strong>of</strong> these methods must be careful in order to avoid experimental errors.<br />

Barge, Edward G and Cathy L Cripps. Plant Sciences and Plant Pathology Department,<br />

Montana State University, Bozeman, MT 59715. Systematics <strong>of</strong> Lactarius<br />

in the Rocky Mountain alpine zone<br />

Lactarius is an important ectomycorrhizal genus in the Arctic-Alpine<br />

Biome where it associates primarily with Salix and Betula species. The Arctic-<br />

Alpine Biome covers roughly 8% <strong>of</strong> the earth’s land and in the Rocky Mountains<br />

<strong>of</strong> North <strong>America</strong> the alpine is comprised <strong>of</strong> scattered “islands” above timberline<br />

in mountainous areas. Beginning in 1999 Cripps, Horak and others surveyed arctic-alpine<br />

macromycete distributions in the Rocky Mountains <strong>of</strong> Montana, Colorado<br />

and Wyoming. Studies <strong>of</strong> the ectomycorrhizal fungi present in arctic-alpine<br />

areas are <strong>of</strong> importance as climate change impacts this biome, which includes the<br />

range expansion <strong>of</strong> Salix species, a key alpine ectomycorrhizal phytobiont. This<br />

study focuses on the systematics <strong>of</strong> Lactarius in alpine areas <strong>of</strong> the Rocky Mountains.<br />

Macromorphological descriptions made at the time <strong>of</strong> collection, microscopic<br />

examination <strong>of</strong> dried material and phylogenetic analysis (rough) have contributed<br />

to identification. Drawings <strong>of</strong> spores, pleuromacrocystidia and<br />

cheilomacrocystidia were completed for each species using a Leica drawing tube<br />

and are complemented with SEM photographs for each species. DNA from upwards<br />

<strong>of</strong> 48 collections was successfully extracted and the ITS region amplified<br />

using primers ITS1-F and ITS4. Preliminary analysis shows morphologically<br />

identified species grouping together nicely. Thus far, the study has resulted in<br />

identification <strong>of</strong> six alpine Lactarius species: L. glyciosmus, L. lanceolatus, L.<br />

nanus, L. pseudouvidus, L. repraesentaneus and L. salicis-reticulatae. Other than<br />

a preliminary report, most are first documentations for these species in alpine<br />

Continued on following page


habitats south <strong>of</strong> the Canadian border. Host patterns are possible among the<br />

species that occur with dwarf and shrub Salix and one species seems to occur only<br />

with Betula. These species are known from many arctic-alpine habitats throughout<br />

the world in places such as Greenland, Iceland, the Alps, Svalbard and Scandinavia<br />

and are now documented for the Rocky Mountains <strong>of</strong> Montana,<br />

Wyoming and Colorado.<br />

Baroni, Timothy J 1 , Ana Esperanza Franco-Molano 2 , Marcelo Betancur 2 , and<br />

Tatiana Sanjuan 2 . 1 Department <strong>of</strong> Biological Sciences, P. O. Box 2000, State<br />

University <strong>of</strong> New York - College at Cortland, Cortland, NY 13045, USA, 2 Laboratorio<br />

de Taxonomía y Ecología de Hongos, Instituto de Biología, Universidad<br />

de Antioquia, A.A. 1226, Medellín, Colombia. New species <strong>of</strong> agarics from the<br />

Páramo region in Colombia<br />

In May <strong>of</strong> 2011 a one-day field excursion was made to the Páramo<br />

ecosystem (Luteyn, 1999) in the department <strong>of</strong> Tolima, Colombia, as part <strong>of</strong> a<br />

longer research excursion to the cloud forest mountainous regions <strong>of</strong> south central<br />

Colombia. Fifteen collections <strong>of</strong> mostly agarics were made in the Páramo,<br />

several growing directly on decaying leaves <strong>of</strong> standing Espeletia. Two <strong>of</strong> these<br />

collections are clearly new species, one a Melanotus, the other a Hypsizygus.<br />

These new taxa will be illustrated and the other collections found on that excursion<br />

will also be displayed and discussed. This isolated area with its distinctive<br />

vegetation clearly deserves further exploration over longer time periods as it appears<br />

to be under explored.<br />

Bartnicki-Garcia, Salomon. Dept. Microbiology, CICESE, Ensenada Center for<br />

Research & Higher Education, Baja California, Mexico. Lysine pathways and<br />

cell wall chemistry revealed the existence <strong>of</strong> two evolutionary lines within the<br />

Kingdom Fungi<br />

Henry Vogel’s discovery (1960’s) splitting fungi into two camps based on<br />

lysine biosynthesis provided the first and strongest piece <strong>of</strong> evidence for the existence<br />

<strong>of</strong> two evolutionary lines within Fungi. By synthesizing lysine via the common<br />

route characteristic <strong>of</strong> bacteria, algae and plants, diaminopimelic acid path<br />

(DAP), Oomycetes and Hypochytridomycetes proved to be distinct from most<br />

Fungi, namely Chytridiomycetes, Zygomycetes, Ascomycetes and Basidiomycetes<br />

which synthesize lysine via aminoadipic acid pathway (AAA), a<br />

unique route found almost exclusively in fungi. Concurrent studies on cell wall<br />

chemistry <strong>of</strong> various fungi revealed drastic differences that correlated with lysine<br />

pathways, thus DAP fungi became members <strong>of</strong> the Cellulosic line, and AAA<br />

fungi <strong>of</strong> the Chitinous line. Later, molecular phylogeny convincingly confirmed<br />

the evolutionary schism among organisms traditionally called fungi but it also<br />

placed Mycology in an existential dilemma: either preserve phylogenetic purity<br />

by removing a most important group <strong>of</strong> organisms, the cellulosic fungi, from the<br />

kingdom or continue maintaining both cellulosic and chitinous fungi within the<br />

kingdom boundaries. While the first alternative has become widely adopted, it has<br />

serious negative repercussions. The second alternative would seem more practical<br />

and beneficial to the science <strong>of</strong> Mycology and to other fields where cellulosic<br />

fungi are <strong>of</strong> great importance. It benefits nobody to exclude Oomycetes from the<br />

real world <strong>of</strong> fungi namely, publications, review articles, congresses, classrooms,<br />

textbooks, and all sorts <strong>of</strong> academic and research opportunities focused on Fungi.<br />

I advocate setting aside the issue <strong>of</strong> phylogenetic purity, and declare the Kingdom<br />

Fungi as made <strong>of</strong> two separate evolutionary lines each giving rise to individuals<br />

different in biochemical details but displaying common physiologic and morphologic<br />

features... and please let’s not refer to a classic fungus <strong>of</strong> the genus Phytophthora<br />

as a “protista-like” organism.<br />

Bates, Scott T 1 , J Gregory Caporaso 2 , D Lee Taylor 3 , and Noah Fierer 1 . 1 Cooperative<br />

Institute for Research in Environmental Sciences, University <strong>of</strong> Colorado,<br />

Boulder, CO, USA, 2 Department <strong>of</strong> Computer Science, Center for Micro-<br />

3<br />

bial Genetics and Genomics, Northern Arizona University, Flagstaff, AZ, USA,<br />

Institute <strong>of</strong> Arctic Biology and Department <strong>of</strong> Biology and Wildlife, University<br />

<strong>of</strong> Alaska Fairbanks, Fairbanks, AK, USA. High-throughput sequencing to explore<br />

the biogeography <strong>of</strong> soil fungi and the use <strong>of</strong> network analyses to examine<br />

fungal-bacterial interactions<br />

High-throughput DNA sequencing technologies have paved the way for<br />

new analytical approaches that move beyond basic descriptive inventories <strong>of</strong> microbial<br />

communities. With the aim <strong>of</strong> deciphering the structure <strong>of</strong> complex microbial<br />

communities across spatial or temporal gradients, network analysis allows<br />

for the examination <strong>of</strong> potential interactions between microbial taxa in natural systems<br />

using large environmental sequence datasets. Here, we sampled soils from<br />

numerous sites across North and South <strong>America</strong>, and used fungal ITS and bacterial<br />

16S rRNA gene sequencing with Illumina technology to recover millions <strong>of</strong><br />

environmental sequences from our sample sites. In addition to examining biogeography<br />

for soil fungi, we also applied network analysis approaches to identify<br />

significant co-occurrence patterns between fungi and bacteria on the basis OTU<br />

abundance and occupancy. We outline factors that influence broad-scale biogeographical<br />

patterns for fungi in terrestrial systems, and describe the topology <strong>of</strong> the<br />

resulting fungal/bacterial co-occurrence network. Finally, we discuss the utility <strong>of</strong><br />

the network approach for exploring inter-taxon correlations to gain a more integrated<br />

understanding <strong>of</strong> microbial community structure and the ecological rules<br />

guiding community assembly.<br />

Beard, Charles E 1 and Abdullah Inci 2 . 1 Clemson University, School <strong>of</strong> Agricultural,<br />

Forest, and Environmental Sciences, 114 Long Hall, Clemson, SC,<br />

29<strong>63</strong>4, USA, 2 Ericyes University, Faculty <strong>of</strong> Veterinary Medicine, Parasitology<br />

Department, Kayseri, Turkey. Potential <strong>of</strong> the trichomycete fungus Harpella<br />

melusinsae to inhabit the midgut <strong>of</strong> mosquito hosts.<br />

Trichomycete fungi are symbiotic inhabitants <strong>of</strong> the guts <strong>of</strong> aquatic<br />

Diptera. The midguts and hindguts <strong>of</strong> Chironomidae (midges) and Simuliidae<br />

(black flies) have trichomycetes, whereas only the hindguts <strong>of</strong> Culicidae (mosquitoes)<br />

have trichomycetes. We asked why no trichomycetes use the midguts <strong>of</strong><br />

mosquitoes even though the midguts are similar among the three families. We hypothesized<br />

four broadly defined limitations that prevent trichomycetes from occupying<br />

mosquito midguts: behavioral, structural, physiological, or host habitat.<br />

The first three limitations were investigated by exposing mosquitoes to Harpella<br />

melusiane, which normally grows only in black fly larval midguts. Larval black<br />

flies from the field sites were allowed to feed in glass jars and shed frass that contained<br />

trichomycete spores. Mosquito larvae were then exposed to the water containing<br />

shed Harpella melusinae trichospores. They were allowed to feed on the<br />

spores, and then assayed for colonization. We found that Harpella melusinae<br />

grew in the mosquito midguts (prevalence up to 66%). These results suggest that<br />

mosquito behavior, gut structure, or physiological limitations do not explain the<br />

lack <strong>of</strong> trichomycete fungi in the midguts <strong>of</strong> field-collected mosquitoes. We suggest<br />

that host-habitat limitations are probably important in limiting trichomycete<br />

colonization <strong>of</strong> mosquito midguts.<br />

Beaulieu, Wesley T 1 , Wittaya Kaonongbua 1 , Daniel G Panaccione 2 , and Keith<br />

Clay 1 . 1 2<br />

Department <strong>of</strong> Biology, Indiana University, Bloomington, IN 47405,<br />

Division <strong>of</strong> Plant and Soil Sciences, West Virginia University, Morgantown,<br />

WV 26506. Molecular, chemical and morphological diversity <strong>of</strong> Periglandula,<br />

clavicipitaceous symbionts <strong>of</strong> Convolvulaceae (Morning Glories)<br />

Fungal endophytes in the Clavicipitaceae (Ascomycota: Hypocreales)<br />

have received considerable attention for production <strong>of</strong> bioactive ergot alkaloids in<br />

associations with monocotyledonous plants, primarily economically important<br />

grasses. The source <strong>of</strong> ergot alkaloids in the dicotyledonous Convolvulaceae<br />

(morning glories) was unknown until the description <strong>of</strong> two clavicipitaceous<br />

fungi, Periglandula ipomoeae and P. turbinae, which form associations with different<br />

host species in the Convolvulaceae. There are dozens <strong>of</strong> Convolvulaceae<br />

known to contain ergot alkaloids, with potentially hundreds more, suggesting<br />

each harbors a clavicipitaceous symbiont responsible for alkaloid production.<br />

Here we report on a molecular, chemical and morphological survey <strong>of</strong> six species<br />

<strong>of</strong> Ipomoea (Convolvulaceae) which contain ergot alkaloids for the presence <strong>of</strong><br />

clavicipitaceous fungal symbionts. All six host species exhibited characteristic<br />

epiphytic mycelia on adaxial surfaces <strong>of</strong> young leaves similar to those <strong>of</strong> described<br />

Periglandula (visualized with SEM) and contained unique pr<strong>of</strong>iles <strong>of</strong><br />

ergot alkaloids in seeds (measured with HPLC). There were considerable intraspecific<br />

differences in the density <strong>of</strong> epiphytic mycelia on young leaves and the<br />

presence and composition <strong>of</strong> ergot alkaloids in foliage. Using PCR and molecular<br />

cloning we sequenced two loci from fungi derived from each host: the internal<br />

transcribed spacer (ITS) region and the dimethylallyl tryptophan synthase gene<br />

(dmaW), which codes for the enzyme that catalyzes the first step in ergot alkaloid<br />

biosynthesis. In phylogenetic analyses, fungal ITS sequences from each host<br />

species were quite similar except for those from the two most divergent hosts (I.<br />

hildebrandtii and I. amnicola) and together formed a clade with described<br />

Periglandula spp. Sequences <strong>of</strong> dmaW were more variable, and sequences from<br />

each fungus formed a distinct clade, suggesting they are unique species, except for<br />

fungi from the two most closely related species (I. gracilis and I. muelleri), which<br />

grouped together. These results suggest there are many species <strong>of</strong> Periglandula<br />

beyond the two described.<br />

Benitez, Maria S 1 , Michelle H Hersh 2 , Brantlee S Richter 3 , Rytas Vilgalys 4 ,<br />

and James S Clark 1 . 1 Nicholas School <strong>of</strong> the Environment, Duke University,<br />

Durham, NC, 2 Program in Biology, Bard College and Cary Institute <strong>of</strong> Ecosystem<br />

Studies, Annandale-on-Hudson, NY, 3 Plant Pathology, University <strong>of</strong> Florida,<br />

Gainesville, FL, 4 Department <strong>of</strong> Biology, Duke University, Durham, NC. Endophyte<br />

and pathogen communities <strong>of</strong> symptomatic and asymptomatic<br />

seedlings <strong>of</strong> tree species from a temperate forest<br />

Fungal pathogens can impact plant community composition through negative<br />

density dependence regulation. Pathogen identification is required to quantify<br />

their effects on plant host species and diversity. To identify potential seedling<br />

pathogens we compared fungal communities associated with symptomatic and<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 9


asymptomatic seedlings from seven plant species sampled at the Duke Forest<br />

(NC) on multiple years. Surface sterilized plant tissue was used for fungal pure<br />

culture isolation and total DNA extraction for community analysis. Taxonomic<br />

identification <strong>of</strong> pure cultures was performed based on sequence information <strong>of</strong><br />

the internal transcribed spacer region (ITS). A collection <strong>of</strong> over 1300 pure cultures,<br />

classified into 316 fungal taxa (97% similarity at the ITS region), was generated<br />

from 320 seedlings. The most abundant fungal groups present in symptomatic<br />

seedlings differed from those most commonly isolated from asymptomatic<br />

seedlings and <strong>of</strong>ten represent known plant pathogenic groups. In addition, the<br />

most abundant pathogen taxa were isolated from five <strong>of</strong> the seven host species<br />

studied, and were isolated at lower frequency from asymptomatic seedlings. The<br />

most abundant isolates from asymptomatic seedlings most closely matched samples<br />

from GenBank with no clear taxonomic identification. For individual<br />

seedlings studied, the fungal community appears to be simple at this stage, dominated<br />

by one or two taxa. At the isolate level, no host-specificity was observed;<br />

however, preliminary results from culture-independent analyses reveal community<br />

structure separation according to host species. The implications <strong>of</strong> host specificity<br />

<strong>of</strong> fungal pathogens in forest seedlings are discussed.<br />

Berbee, Mary L 1 , Satoshi Sekimoto 1 , Ludovic LeRenard 1 , Joseph W Spatafora<br />

2 , and AFTOL2 Working Group 1 . 1 Dept. <strong>of</strong> Botany, University <strong>of</strong> British Columbia,<br />

Vancouver BC Canada, 2 Dept. <strong>of</strong> Botany and Plant Pathology, Oregon<br />

State University, Corvallis OR USA. Becoming A Fungus: Comparative Phylogenetic<br />

Studies <strong>of</strong> Evolution <strong>of</strong> Absorptive Nutrition<br />

Most fungi gain nutrients by secreting digestive enzymes from their hyphae<br />

into the surrounding matrix, and then absorbing the nutrients that diffuse<br />

back. We are exploring data from our community sequencing proposal to the US<br />

Joint Genome Institute to better understand the metabolic capabilities <strong>of</strong> early<br />

fungi. With new genome sequences from early-diverging fungi including aquatic<br />

chytrids and terrestrial zygomycetes, it becomes possible to compare secreted enzymes<br />

across phyla and to speculate on the evolutionary origins <strong>of</strong> their underlying<br />

genes. The enzymes that fungi secrete to break down cellulose or proteins<br />

have paralogs that are not secreted but function instead in housekeeping within the<br />

cell, in modifying polysaccharides or in recycling proteins. Duplication <strong>of</strong> the<br />

housekeeping genes, modification for secretion, and addition <strong>of</strong> domains for binding<br />

to particular substrates can result in enzymes adapted for extracellular function.<br />

The Ascomycota and Basidiomycota share secreted enzymes with functions<br />

including breaking down cellulose. The chytrids and zygomycetes also have the<br />

housekeeping homologs to the secreted enzymes. However, true orthologs to the<br />

secreted enzymes <strong>of</strong> the Ascomycota and Basidiomycota seem to be rare among<br />

the early diverging clades. Assuming that it was the accumulation <strong>of</strong> plant biomass<br />

on land that selected for the maintenance <strong>of</strong> the enzymes <strong>of</strong> the Ascomycota<br />

and Basidiomycota, the absence <strong>of</strong> conserved secreted enzymes across other<br />

phyla could be explained if other fungi diverged too early to encounter land plants<br />

as a readily available source <strong>of</strong> nutrients.<br />

Birkebak, Joshua M and Brandon Matheny. Department <strong>of</strong> Ecology and Evolutionary<br />

Biology, University <strong>of</strong> Tennessee, 332 Hesler Biology Building<br />

Knoxville, TN 37996-1610. A systematic revision <strong>of</strong> Clavariaceae (Agaricales)<br />

from the Pacific Northwest<br />

The diversity <strong>of</strong> Clavariaceae has been underestimated in the Pacific<br />

Northwest <strong>of</strong> North <strong>America</strong> and thorough taxonomic revisions are much needed.<br />

Molecular methods <strong>of</strong> species recognition have not been extensively used in<br />

the Clavariaceae and are found to elucidate species level diversity previously<br />

unidentified. Several species complexes exhibiting high morphological and sequence<br />

diversity have been observed. Morphological differentiation <strong>of</strong> some molecularly<br />

distinctive species is not yet possible, and cryptic species may be present.<br />

A monographic revision <strong>of</strong> the Clavariaceae in the Pacific Northwestern North<br />

<strong>America</strong> with taxonomic keys, illustrations and descriptions is being compiled<br />

and will be published pending further investigation. At present, twenty-eight<br />

species in the family have been identified based on herbarium and recent field collections.<br />

Ten <strong>of</strong> these are tentatively considered undescribed. Overall, the Pacific<br />

Northwest is represented by two species <strong>of</strong> Camarophyllopsis (one new), eight<br />

species <strong>of</strong> Clavaria (one new), one species <strong>of</strong> Clavicorona, four species <strong>of</strong><br />

Clavulinopsis (two new), four species <strong>of</strong> Mucronella, and nine species <strong>of</strong> Ramariopsis<br />

(six new).<br />

Bittleston, Leonora S and Anne Pringle. Harvard University,16 Divinity Avenue<br />

Cambridge, MA 02138. The insect and yeast communities <strong>of</strong> carnivorous<br />

pitcher plants<br />

Carnivorous pitcher plants are models for food web dynamics. The Northern<br />

pitcher plant, Sarracenia purpurea, has modified leaves, or pitchers, which<br />

are sterile until they open. Soon after opening the pitchers are filled with rainwater,<br />

and accumulate a diverse community <strong>of</strong> organisms. The microbes in pitchers<br />

are still relatively unknown, although recent studies have shown that a keystone<br />

predator, the mosquito Wyeomyia smithii, controls bacterial diversity. Numerous<br />

10 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

yeasts have been found within the pitchers, and they are different from those present<br />

in the surrounding bog water. We examined how insects affected the diversity<br />

and abundance <strong>of</strong> yeasts in S. purpurea pitcher plants. Insect exclusion with<br />

gauze coverings successfully excluded W. smithii, and the abundance <strong>of</strong> yeasts<br />

was positively correlated with insect counts. The gauze treatment did not reduce<br />

the abundance or diversity <strong>of</strong> yeasts in the pitchers, suggesting that insects introduce,<br />

attract, or promote the growth <strong>of</strong> yeasts. Additionally, we found that one<br />

commonly associated yeast, Candida globosa, is present internally in surfacesterilized<br />

adult W. smithii, indicating that the pitcher plant mosquito acts as a vector<br />

and may have a more complex association with this yeast. Convergently<br />

evolved pitcher plants in the genus Nepenthes exist in Southeast Asia, and our preliminary<br />

research show that there is also convergent community assembly <strong>of</strong> insect<br />

and arachnid associates between Nepenthes and Sarracenia. These results<br />

may extend to the communities <strong>of</strong> yeasts found within similar pitcher habitats on<br />

opposite sides <strong>of</strong> the planet.<br />

Blair, Jaime E and Nahill H Matari. Department <strong>of</strong> Biology, Franklin & Marshall<br />

College, Lancaster, PA 17603. A timescale for Oomycete evolution estimated<br />

from conserved regulators <strong>of</strong> gene expression<br />

The fungal-like oomycetes are ubiquitous in nature, occupying niches in<br />

marine, freshwater, and terrestrial ecosystems. While the diversity <strong>of</strong> saprophytic<br />

oomycetes is most certainly underestimated, this group is primarily known for the<br />

important plant and animal pathogens it contains. The oldest accepted fossil evidence<br />

<strong>of</strong> biotrophic oomycetes associated with vascular plants comes from the<br />

Early Devonian Rhynie chert (approx 400 Ma); however, the affinity <strong>of</strong> fossil<br />

oomycetes to modern lineages remains unclear. Currently, genomic resources<br />

exist for a number <strong>of</strong> oomycete species, as well as close relatives within the Stramenopiles<br />

(diatoms, brown algae). The goal <strong>of</strong> this project was to use a comparative<br />

genomics approach to identify conserved regulators <strong>of</strong> gene expression<br />

within each <strong>of</strong> these genomes, and use Bayesian techniques to co-estimate phylogeny<br />

and divergence times. We focused primarily on genes involved in key<br />

processes such as RNA silencing and chromatin modification, as well as core eukaryotic<br />

transcription factors. Oomycete genomes do contain the necessary complement<br />

<strong>of</strong> genes for RNA silencing, including orthologs <strong>of</strong> Dicer, Argonaut,<br />

dsRNA-binding proteins, and an RNA-dependent RNA polymerase; other researchers<br />

have shown that these genes are expressed during various life stages in<br />

Phytophthora infestans. Oomycetes also possess a full array <strong>of</strong> genes for histone<br />

modification, including deacetylases, acetyltransferases, methyltransferases, and<br />

chromodomain-containing proteins. Surprisingly, no homologs to canonical eukaryotic<br />

DNA methyltransferases could be identified in the genomes <strong>of</strong><br />

oomycetes or in the outgroups. Bayesian divergence times were estimated from a<br />

dataset <strong>of</strong> 53 proteins using conservative calibrations from the diatoms and the<br />

Rhynie chert oomycetes. The sensitivity <strong>of</strong> estimated divergence times to model<br />

specification on the prior parameters will also be discussed.<br />

Blair, Jaime E, Lauren C<strong>of</strong>fua, Alison Greidinger, Amy Chabitnoy, and Lauren<br />

Cook. Department <strong>of</strong> Biology, Franklin & Marshall College, Lancaster, PA<br />

17603. Measuring oomycete biodiversity in aquatic, forest, and agricultural<br />

ecosystems: culture-based and metagenomic approaches<br />

The fungal-like Oomycota have long been studied by mycologists given<br />

their osmotrophic lifestyle and habit <strong>of</strong> hyphal growth. Many species <strong>of</strong><br />

oomycetes are devastating pathogens, such as the notorious potato late blight<br />

agent, Phytophthora infestans, and the cause <strong>of</strong> rare human pythiosis, Pythium insidiosum;<br />

it is not surprising then that most research has focused on host-pathogen<br />

interactions, with relatively little study concerning the natural diversity <strong>of</strong> saprotrophic<br />

species. This knowledge gap has important ramifications on our estimates<br />

<strong>of</strong> oomycete biodiversity, as well as on our understanding <strong>of</strong> the evolutionary history<br />

<strong>of</strong> this important group. The goal <strong>of</strong> this study is to estimate and compare<br />

oomycete biodiversity from several distinct habitats, including aquatic environments,<br />

undisturbed forest soils, and highly managed agricultural settings. Our ongoing<br />

culture-based surveys have relied on various baiting techniques and have<br />

revealed a high level <strong>of</strong> diversity in aquatic and agricultural environments. However,<br />

it has been shown that baiting methods can lead to biased estimates <strong>of</strong> biodiversity<br />

as they tend to favor fast-growing organisms and those producing motile<br />

zoospores. We have therefore developed a complementary metagenomic approach<br />

using massively parallel pyrosequencing to more thoroughly sample<br />

species diversity from the different environments. We are currently verifying two<br />

markers for analysis, the mitochondrial cytochrome c oxidase subunit 1 (cox1)<br />

locus and the nuclear ribosomal RNA internal transcribed spacers (ITS) and large<br />

subunit (LSU) region; previous studies have shown that these loci are able to discriminate<br />

among closely related species, and provide phylogenetic signal at a<br />

number <strong>of</strong> taxonomic levels. We expect that this combination <strong>of</strong> culture-based<br />

and sequence-based identification methods will enhance our understanding <strong>of</strong><br />

oomycete ecology and evolution, and perhaps give us more insight into the roles<br />

<strong>of</strong> certain species in the outbreak <strong>of</strong> disease.<br />

Continued on following page


Boddy, Lynne. Cardiff School <strong>of</strong> Biosciences, Cardiff University, Cardiff CF10<br />

3AS, UK. Interactions among saprotrophic fungi and invertebrate grazers<br />

under climate change<br />

Interspecific interactions between fungi drive changes in fungal community<br />

structure. The progress and final outcome <strong>of</strong> these interactions, i.e. deadlock<br />

or replacement, are affected by the abiotic environment, changes in temperature,<br />

water potential and CO2 regime sometimes completely reversing outcomes. Similarly,<br />

invertebrate grazers can exert selective pressures on fungal decomposer<br />

communities in soil by feeding selectively on certain fungi and reversing the outcomes<br />

<strong>of</strong> competitive interactions. For example, by feeding selectively on the<br />

cord-forming fungus Resinicium bicolor, isopods prevented the competitive exclusion<br />

<strong>of</strong> Hypholoma fasciculare and Phanerochaete velutina in soil and wood.<br />

Nematode populations also reversed the outcomes <strong>of</strong> competitive interactions by<br />

stimulating growth <strong>of</strong> less competitive fungi. Composition <strong>of</strong> fungal communities<br />

affects decomposition rate, but so also can fungal-fungal and fungal-invertebrate<br />

interactions themselves stimulate or inhibit decomposition and extracellular enzyme<br />

production. Climate change has the potential to alter the activity <strong>of</strong>, and interactions<br />

between, saprotrophic fungi and soil invertebrate grazers, with implications<br />

for decomposer community composition, decomposition, ecosystem<br />

regulation and carbon feedback.<br />

Bojantchev, Dimitar. 345 Shipwatch Ln, Hercules, CA 94547. Vernal fruiting<br />

Cortinarii from the mountains <strong>of</strong> California and the Pacific Northwest<br />

The mountains <strong>of</strong> California and Western North <strong>America</strong> feature a diverse<br />

set <strong>of</strong> Cortinarius species that fruit in the spring, during, or shortly after<br />

snowmelt. This vernal-fruiting phenomenon is much more prevalent in the <strong>America</strong>n<br />

West than in Europe. Based on several years <strong>of</strong> intense collecting in California<br />

and adjacent states, utilizing morphological and molecular methods, we have<br />

become familiar with 26 Cortinarius species that fruit in the spring and early summer.<br />

Their subgeneric breakdown is 16 species in subg. Telamonia, 8 in subg.<br />

Phlegmacium (including 2 hypogeous forms), 2 in subg. Dermocybe, 1 in sect.<br />

Leprocybe and 1 in sect. Anomali. About one third <strong>of</strong> these species are very rare<br />

and so far represented by single collections only. Every season we encounter<br />

species that we had not collected before. Almost all <strong>of</strong> these are new to science<br />

and we are in the process <strong>of</strong> describing them. Some <strong>of</strong> these vernal fruiting Cortinarii<br />

have a broad distribution throughout the West, while others seem to exhibit<br />

a degree <strong>of</strong> endemism, as so far they are known only from California. Most <strong>of</strong><br />

these vernal fruiting species span the clades within genus Cortinarius, except for<br />

one group <strong>of</strong> Telamonia featuring UV active universal veil that form a well-defined<br />

clade (/colymbadinus). Only two <strong>of</strong> the vernal species are known to fruit in<br />

the fall. In this session will review the current state <strong>of</strong> the studies and present the<br />

challenges and directions <strong>of</strong> future research.<br />

Bonito, Gregory 1 , Christopher Schadt 2 , Gerald (Jerry) Tuskan 2 , Mitchel Doktycz<br />

2 , and Rytas Vilgalys 1 . 1 2<br />

Biology Department, Duke University, Durham NC,<br />

Oak Ridge National Laboratory, Department <strong>of</strong> Energy, Oak Ridge TN 37831.<br />

Populus species, genotype, and soil inoculum influences on the assemblage <strong>of</strong><br />

rhizospheric fungal and bacterial communities<br />

Populus deltoides is a common riparian tree species in southeastern North<br />

<strong>America</strong>. Populus forms root associations with both arbuscular and ectomycorrhizal<br />

fungi and also with bacterial and fungal endophytes. To address the influence<br />

<strong>of</strong> edaphic and genotypic factors on the structuring <strong>of</strong> rhizospheric assemblages,<br />

we carried out a series <strong>of</strong> trap-plant experiments using rooted cuttings <strong>of</strong><br />

various Populus genotypes. Two other tree species (Quercus phellos and Pinus<br />

taeda) were included for comparison. Plants were grown in field soils in growth<br />

chambers. We used 454 multiplex amplicon pyrosequencing to characterize fungal<br />

and bacterial root communities. Specifically, we were interested in the effects<br />

<strong>of</strong> host species, P. deltoides genotype, and soil inoculum on the structuring <strong>of</strong> rhizospheric<br />

communities. Although species richness <strong>of</strong> endophytic root fungi was<br />

higher for Populus than for oak or pine, Populus hosted fewer ectomycorrhizal<br />

taxa. Total richness <strong>of</strong> root associated bacteria taxa (283) was also significantly<br />

greater for Populus than for oak (175-f;184-b) or pine (157-f;185-b) and Populus<br />

was characterized by a higher relative abundance <strong>of</strong> Actinobacteriales and Sphingobacteriales,<br />

and a lower abundance <strong>of</strong> Rhizobiales and Burkholderiales. There<br />

were minor responses <strong>of</strong> bacterial and fungal communities to Populus genotype,<br />

but the Populus hybrid associated with a wider variety <strong>of</strong> fungi and a greater frequency<br />

and relative abundance <strong>of</strong> ectomycorrhizal taxa (e.g. Inocybe, Tomentella,<br />

Hebeloma). Arbuscular mycorrhizal fungi belonging to the Glomerales and<br />

Paraglomales were present in all Populus genotypes but sequences belonging to<br />

the Diversisporales were only detected in the hybrid Populus genotype. Soils had<br />

a large impact on microbial communities, and abundant species in some soils<br />

were absent in others. In conclusion, rhizospheric microbiota <strong>of</strong> P. deltoides is diverse<br />

and unique from other tree species, and appears to be structured both by the<br />

microbial inoculum available in soils and, to a lesser extent, by plant genotype.<br />

Bonito, Gregory1 , Matthew Smith, Michael Nowak, Rosanne Healy, Gonzalo<br />

Guevara, Efren Cázares, Akihiko Kinoshita, Eduardo Nouhra, Laura Domínguez,<br />

Leho Tedersoo, Claude Murat, Yun Wang, Baldomero Arroyo Moreno, Donald<br />

Pfister, Kazuhide Nara, Alessandra Zambonelli, James Trappe, and Rytas Vilgalys.<br />

1Deparment <strong>of</strong> Biology, 125 Science Drive, Duke University, Durham NC<br />

27708, USA. Phylogeography and diversification <strong>of</strong> Truffles in the Tuberaceae<br />

Truffles have evolved from epigeous (above-ground) ancestors in nearly<br />

every major lineage <strong>of</strong> fleshy fungi. Because accelerated rates <strong>of</strong> morphological<br />

evolution accompany the transition to the truffle form, closely related epigeous<br />

ancestors remain unknown for most truffle lineages. This is the case for the quintessential<br />

truffle genus Tuber, which includes species with socio-economic importance<br />

and esteemed culinary attributes. Ecologically, Tuber spp. form obligate<br />

mycorrhizal symbioses with a diverse species <strong>of</strong> plant hosts, including pines,<br />

oaks, poplars, orchids, and commercially important trees such as hazelnut and<br />

pecan. Unfortunately, limited geographic sampling and inconclusive phylogenetic<br />

relationships have obscured our understanding <strong>of</strong> their origin, biogeography,<br />

and diversification. To address this problem, we present a global sampling <strong>of</strong> Tuberaceae<br />

based on DNA sequence data from four loci for phylogenetic inference<br />

and molecular dating. Our well-resolved Tuberaceae phylogeny shows high levels<br />

<strong>of</strong> regional and continental endemism. We also identify a previously unknown<br />

epigeous member <strong>of</strong> the Tuberaceae - the South <strong>America</strong>n cup-fungus Nothojafnea<br />

thaxteri (Cash) Gamundí. Phylogenetic resolution was further improved<br />

through the inclusion <strong>of</strong> a previously unrecognized Southern Hemisphere sister<br />

group <strong>of</strong> the Tuberaceae. This morphologically diverse assemblage <strong>of</strong> fungi includes<br />

truffle forms endemic to Australia and South <strong>America</strong> (e.g. Gymnohydnotrya<br />

spp.) and non-truffle species. The Southern hemisphere taxa appear to<br />

have diverged more recently than the Northern Hemisphere lineages. Our analysis<br />

<strong>of</strong> the Tuberaceae suggests that Tuber evolved from an epigeous ancestor in<br />

close association with angiosperm hosts. Molecular dating estimates <strong>of</strong> ~134 million<br />

years ago for Tuberaceae are in accord with this hypothesis. Intra-continental<br />

diversification, limited long-distance dispersal, and ecological adaptations<br />

have driven evolution and Tuberaceae biodiversity.<br />

Boonmee, Saranyaphat and Amy Y Rossman. Systematic Mycology and Microbiology<br />

Laboratory, USDA-ARS, Beltsville, Maryland 20705, USA. Tropical<br />

species <strong>of</strong> Tubeufiaceae (Dothideomycetes) in northern Thailand<br />

Fifteen species <strong>of</strong> Tubeufiaceae were found on decaying wood in northern<br />

Thailand. Genera in this family are characterized by pseudothecial ascomata<br />

that are superficial, light brown, dark brown to black, <strong>of</strong>ten smooth. The bitunicate<br />

asci are cylindrical to broadly clavate and contain eight ascospores that are<br />

filiform, cylindrical to narrowly fusiform, tapering towards rounded to sub-acute<br />

ends, multiseptate, hyaline, and pale yellow or brown. Some <strong>of</strong> these fungi produce<br />

a hyphomycetous asexual state in culture such as Chlamydotubeufia khunkornensis<br />

with chlamydospores, Thaxteriella inthanonensis with helicoma-like<br />

conidia, and Thaxteriellopsis lignicola with helicomyces-like conidia. The relationship<br />

between sexual and asexual states is poorly known especially for species<br />

from Thailand. Several <strong>of</strong> these taxa appear to be new to science.<br />

Borovicka, Jan 1,2 , Martin Mihaljevic 3 , and Milan Gryndler 4 . 1 Academy <strong>of</strong> Sciences<br />

<strong>of</strong> the Czech Republic, Nuclear Physics Institute, v.v.i., Department <strong>of</strong> Nuclear<br />

2<br />

Spectroscopy, Rez 130, CZ-250 68 Rez near Prague, Czech Republic,<br />

Academy <strong>of</strong> Sciences <strong>of</strong> the Czech Republic, Institute <strong>of</strong> Geology, v.v.i., Laboratory<br />

<strong>of</strong> Environmental Geology and Geochemistry, Rozvojová 269, CZ-165 00<br />

Prague 6, Czech Republic, 3 Charles University, Faculty <strong>of</strong> Science, Institute <strong>of</strong><br />

Geochemistry, Mineralogy and Mineral Resources, Albertov 6, CZ-12843 Prague<br />

2, Czech Republic, 4 Academy <strong>of</strong> Sciences <strong>of</strong> the Czech Republic, Institute <strong>of</strong> Microbiology,<br />

v.v.i., Laboratory <strong>of</strong> Fungal Biology, Vídenská 1083, CZ-142 20<br />

Prague 4, Czech Republic. Lead isotopic composition <strong>of</strong> macr<strong>of</strong>ungi: possible<br />

applications in fungal ecology<br />

Macr<strong>of</strong>ungi are able to effectively take up various trace elements, including toxic<br />

heavy metals, translocate them through mycelia and deposit them in fruiting bodies.<br />

This ability is species-specific; certain macr<strong>of</strong>ungal species are effective accumulators<br />

<strong>of</strong> such metals as cadmium, zinc, mercury, silver, gold and vanadium.<br />

Despite growing at sites with an arsenic/silver soil concentration at “background”<br />

levels, several macr<strong>of</strong>ungal species have been reported to hyperaccumulate arsenic<br />

or silver. The mechanisms <strong>of</strong> uptake and transport <strong>of</strong> the accumulated elements<br />

are unclear. However, a lead isotopic analysis might represent an efficient<br />

tool for tracing the origin <strong>of</strong> elements in soil pr<strong>of</strong>ile. Lead is a non-essential toxic<br />

metal whose biogeochemical cycle has been affected to a great degree by man and<br />

lead isotopes have been thus introduced as “fingerprints” <strong>of</strong> environmental pollution.<br />

Each source <strong>of</strong> lead can have distinct or sometimes overlapping isotopic ratio<br />

ranges and the isotopic composition <strong>of</strong> lead in soils reflects a mixing <strong>of</strong> these<br />

sources. At sites with contrasting 206 Pb/ 207 Pb and 208 Pb/ 206 Pb ratios in partic-<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 11


ular soil horizons, the isotopic composition <strong>of</strong> macr<strong>of</strong>ungal fruiting bodies might<br />

be a useful tool for tracing lead origin. Our preliminary data from a case study in<br />

Agaricus bernardii have supported this hypothesis. Furthermore, the isotopic<br />

composition <strong>of</strong> lead in saprotrophic macr<strong>of</strong>ungi (Agaricus campestris, A.<br />

bernardii, A. xanthodermus and Leucoagaricus leucothites) <strong>of</strong> various origins has<br />

reflected more or less the expected pollution sources.<br />

Brewer, Marin T and Ashley N Turner. Department <strong>of</strong> Plant Pathology, University<br />

<strong>of</strong> Georgia, Athens, GA 30602. Genetic diversity and reproductive isolation<br />

in Didymella bryoniae, a fungal pathogen <strong>of</strong> cucurbits<br />

We are interested in determining the center <strong>of</strong> diversity, as well as historical<br />

and ongoing migration patterns in the cucurbit pathogen Didymella bryoniae.<br />

This fungus causes gummy stem blight and black rot <strong>of</strong> cucurbits everywhere they<br />

are grown, yet relatively little is known about how it is dispersed. Previous studies<br />

with dominant markers have indicated that this species is diverse and contains<br />

at least two distinct genetic groups. We sequenced four nuclear regions, including<br />

calmodulin (CAL), beta-tubulin (TUB2), chitin synthase 1 (CHS-1), and the internal<br />

transcribed spacer <strong>of</strong> the rDNA (ITS) <strong>of</strong> thirty-four isolates from diverse cucurbit<br />

hosts across the United States (US). Isolates from both genetic groups were<br />

represented in our sample. A total <strong>of</strong> 1415 bp were sequenced, resulting in 41<br />

polymorphic sites and 13 unique haplotypes. A haplotype network constructed<br />

with TCS demonstrated that there are two distinct genetic groups, which is consistent<br />

with previous studies. However, a single isolate collected from a greenhouse<br />

in California was quite distinct from both groups. There is no association<br />

with genetic group and host species <strong>of</strong> origin, which is not surprising because earlier<br />

studies indicated that host specialization is not evident with D. bryoniae. Geographic<br />

associations were detected, however, as one <strong>of</strong> the groups was present<br />

only in the northeastern US whereas the other group was found throughout the<br />

US. There are no detectable morphological differences between the two genetic<br />

groups indicating that these groups represent cryptic species that are reproductively<br />

isolated. Studies on a worldwide collection <strong>of</strong> isolates are currently underway.<br />

Coalescent analyses on this collection will be aimed at understanding migration<br />

patterns, geographic origin, and approximate age <strong>of</strong> the divergence <strong>of</strong> the<br />

genetic groups to elucidate the evolutionary processes contributing to the observed<br />

patterns <strong>of</strong> diversity.<br />

Broders, Kirk 1 , Andre Boraks 1 , Laura Barbison 2 , John Brown 2 , and Greg<br />

Boland 2 . 1 Department <strong>of</strong> Biological Sciences, University <strong>of</strong> New Hampshire,<br />

Durham, NH 03824, 2 School <strong>of</strong> Environmental Sciences, University <strong>of</strong> Guelph,<br />

Guelph, ON N1G 2W1. Invasion biology <strong>of</strong> the butternut canker fungus<br />

Ophiognomonia clavigignenti-juglandacearum<br />

Butternut canker caused by the fungal pathogen Ophiognomonia<br />

clavigignenti-juglandacearum (Oc-j) remains the primary cause for range-wide<br />

mortality <strong>of</strong> butternut trees. The disease was first reported in Wisconsin in 1967,<br />

however the fungus may have been present for several years prior. Several questions<br />

still remain largely unanswered regarding the invasion by Oc-j. These questions<br />

include; how many times was the fungus introduced; where was it introduced;<br />

were more virulent strains introduced; and how was it able to spread so<br />

rapidly within the native butternut population. Therefore, our objective was to<br />

evaluate the invasion biology <strong>of</strong> this fungus including investigations into the ecology,<br />

epidemiology, and population structure <strong>of</strong> Oc-j in North <strong>America</strong>. To complete<br />

these objective flowers, developing seeds, and mature seeds <strong>of</strong> butternut,<br />

heartnut and black walnut were assessed as potential vectors <strong>of</strong> Oc-j; Sixteen isolates<br />

were evaluated for virulence on butternut, heartnut, and black walnut; and<br />

100 isolates <strong>of</strong> Oc-j from across North <strong>America</strong> were used to analyze the population<br />

structure <strong>of</strong> the fungus. Bayesian analyses based on 16 SNP markers revealed<br />

that the Oc-j population in North <strong>America</strong>n is composed <strong>of</strong> four genetically distinct<br />

clonal lineages, suggesting multiple introductions <strong>of</strong> Oc-j, through successive<br />

or simultaneous introductions <strong>of</strong> isolates having differentiated genetic backgrounds.<br />

Isolates <strong>of</strong> Oc-j recovered from heartnut and black walnut caused larger<br />

lesions on all three Juglans species compared to isolates originally recovered from<br />

butternut. Four <strong>of</strong> the five isolates, which caused the largest lesions on butternut<br />

belonged to a single genetic cluster. The pathogenicity data in combination with<br />

geographic and population structure data indicate this fungus was introduced into<br />

North <strong>America</strong> on multiple occasions, and a more virulent clonal lineage was recently<br />

introduced into Minnesota and Wisconsin and subsequently spread<br />

throughout the rest <strong>of</strong> the butternut canker population in North <strong>America</strong>.<br />

Buchanan, Peter K and Peter R Johnston. Landcare Research, Private Bag<br />

92170, Auckland 1142, New Zealand. Conservation <strong>of</strong> Fungi - threat status <strong>of</strong><br />

fungi in New Zealand and globally<br />

Fungi have been included in threat status assessments <strong>of</strong> New Zealand’s<br />

biota since 2002. Assessments have mainly addressed macr<strong>of</strong>ungi, as well as obligate<br />

species <strong>of</strong> fungi on threatened plants. The initial inclusion <strong>of</strong> fungi, and subsequent<br />

need for reassessment <strong>of</strong> threat status, has generated new research initiatives<br />

and raised awareness <strong>of</strong> fungal conservation. Aspects <strong>of</strong> fungal biology such<br />

12 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

as ephemeral reproductive stages add difficulty to threat status assessments. Examples<br />

are presented <strong>of</strong> fungal species listed in the highest (Nationally Critical)<br />

and other threat categories. Over 1,000 species <strong>of</strong> fungi are listed as Data Deficient<br />

due to inadequate distribution data. Fungi have been included in a prioritization<br />

exercise spanning all New Zealand’s threatened taxa that are in decline, to<br />

evaluate methodology, feasibility, and cost <strong>of</strong> long-term recovery plans. Using<br />

molecular techniques, new records <strong>of</strong> ectomycorrhizal fungi listed as Data Deficient<br />

have been discovered by comparing environmental sampling <strong>of</strong> ectomycorrhizas<br />

with DNA sourced from herbarium specimens. Responding to increasing<br />

global awareness <strong>of</strong> the need for fungal conservation, the International <strong>Society</strong> for<br />

Conservation <strong>of</strong> Fungi was formed in 2010, and five Fungal Specialist Groups<br />

have been established under IUCN.<br />

Burchhardt, Kathleen M and Marc A Cubeta. North Carolina State University,<br />

Department <strong>of</strong> Plant Pathology, Raleigh, NC 27695. Population dynamics <strong>of</strong><br />

Monilinia vaccinii-corymbosi in blueberry fields throughout the United<br />

States<br />

Mummy berry disease <strong>of</strong> blueberry (Vaccinium spp.) is caused by the ascomycete<br />

Monilinia vaccinii-corymbosi (Mvc). The fungus produces asexual and<br />

sexual spores that infect blueberry shoots and fruit. While spore dispersal and sexual<br />

reproduction can contribute to the spread and evolution <strong>of</strong> fungal populations,<br />

little is known about the population dynamics <strong>of</strong> Mvc. The primary objective <strong>of</strong><br />

this study was to utilize microsatellite-based genetic markers to examine intraspecific<br />

genetic diversity, population structure, and gene flow among populations<br />

<strong>of</strong> Mvc throughout blueberry growing regions <strong>of</strong> the United States. In this<br />

study, 438 isolates <strong>of</strong> Mvc sampled from 15 blueberry fields in New York, New<br />

Jersey, Massachusetts, Georgia, Mississippi, Oregon, Washington, Michigan, and<br />

North Carolina were screened with 10 polymorphic microsatellite markers. Results<br />

based on analyzing a geographically diverse subsample <strong>of</strong> 55 isolates from<br />

six fields indicate high intraspecific genetic diversity. Forty-eight private alleles<br />

were detected across the 10 loci, providing evidence for population differentiation<br />

and restricted gene flow among fields. Results from population genetic analyses<br />

<strong>of</strong> a more comprehensive dataset <strong>of</strong> 438 isolates will be presented.<br />

Burleigh, J Gordon 1 , Keith Crandall 2 , Karen Cranston 3 , Karl Gude 4 , David S<br />

Hibbett 5 , Mark Holder 6 , Laura A Katz 7 , Richard H Ree 7 , Stephen A Smith 7 ,<br />

Douglas E Soltis 7 , and Tiffani Williams 7 . 1 Department <strong>of</strong> Biology, University <strong>of</strong><br />

Florida, Gainesville, FL 32611, 2 Department <strong>of</strong> Biology, Brigham Young University,<br />

Provo, UT 84602, 3 NESCENT, Durham, NC 27705, 4 College <strong>of</strong> Communication,<br />

5<br />

Arts and Sciences, Michigan State University, E. Lansing<br />

6<br />

MI 45622,<br />

Biology Department, Clark University, Worcester MA 01610, Department <strong>of</strong><br />

Ecology<br />

7<br />

and Evolutionary Biology, University <strong>of</strong> Kansas, Lawrence KS 66045,<br />

and elsewhere. Open Tree <strong>of</strong> Life: Community driven synthesis <strong>of</strong> the Tree<br />

<strong>of</strong> Life<br />

Reconstructing the phylogeny <strong>of</strong> all species has been a grand challenge<br />

ever since Darwin. The scope <strong>of</strong> the problem is immense: current estimates <strong>of</strong> extant<br />

biodiversity range from 1.8 million to 8.7 million species, a large fraction <strong>of</strong><br />

which are Fungi. Much progress has recently been made in resolving the tree, and<br />

systematists continue to generate new phylogenetic knowledge at all depths <strong>of</strong> ancestry.<br />

Nevertheless, despite 150 years <strong>of</strong> effort, 55 AToL projects, and numerous<br />

other funded projects, we lack a comprehensive tree <strong>of</strong> life. Synthesis is currently<br />

inhibited by limits <strong>of</strong> available data, analytical power, and informatics infrastructure.<br />

Perhaps more importantly, it is also limited by a lack <strong>of</strong> compelling<br />

means and incentives for community participation. A comprehensive synthesis<br />

would yield great benefits across the life sciences, especially if it were self-sustaining,<br />

community-driven, and continually updated. We describe a recently funded<br />

project named “Open Tree <strong>of</strong> Life” that aims to establish a community-driven,<br />

continually updated estimate <strong>of</strong> the entire tree, and develop new s<strong>of</strong>tware tools<br />

and new methods for merging and sharing data. Open Tree <strong>of</strong> Life will: 1) within<br />

one year, compile the first comprehensive draft tree <strong>of</strong> life by synthesizing existing<br />

phylogenetic and taxonomic knowledge; 2) enable the community to improve,<br />

annotate, and expand this initial tree; 3) initiate a cultural transformation in<br />

systematics towards pervasive and ingrained practices <strong>of</strong> data sharing; and 4) develop<br />

novel methods for synthetic tree reconstruction. By engaging the systematics<br />

community, including mycologists, our overarching goal is to cultivate ongoing<br />

synthesis on a large scale, in a manner that will transform current cultural<br />

norms in the field.<br />

Bushley, Kathryn E and Joseph W Spatafora. Department <strong>of</strong> Botany and Plant<br />

Pathology, Oregon State University, Corvallis, OR 97330. Convergent and divergent<br />

evolution <strong>of</strong> nonribosomal peptide synthetases<br />

Nonribosomal peptide synthetases (NRPSs) are large multimodular enzymes<br />

which produce small bioactive peptides (NRPs) without the aid <strong>of</strong> ribosomes.<br />

While the function <strong>of</strong> many NRPs remains unknown, it is becoming clear<br />

Continued on following page


that many help shape the ecologies and lifestyles <strong>of</strong> their producing fungi. Production<br />

<strong>of</strong> a host-selective toxin, for example, confers the ability to colonize a specific<br />

host while production <strong>of</strong> antibiotics may increase competitive ability in a particular<br />

habitat. As such, the ability to produce these compounds is likely under<br />

selective pressure. Drawing on examples from across the fungi and focusing on<br />

the order Hypocreales, which consists <strong>of</strong> fungi with diverse ecologies and<br />

lifestyles ranging from plant pathogens to insect/animal pathogens, to fungal<br />

pathogens, we explore correlations between the evolution <strong>of</strong> NRPSs biosynthetic<br />

pathways, their chemical products, and fungal lifestyles. We examine expansions<br />

and contractions <strong>of</strong> phylogenetic clades <strong>of</strong> NRPS adenylation domains correlated<br />

with lifestyle and address evolutionary and genetic mechanisms that may lead to<br />

diversification <strong>of</strong> chemical products. We also address the extent to which these<br />

correlations result from independent expansions versus acquisition by horizontal<br />

transfer or non-vertical evolutionary pathways.<br />

Cafaro, Matias J. Department <strong>of</strong> Biology, University <strong>of</strong> Puerto Rico, Mayaguez,<br />

PR 00681. The protist trichomycete genus Enterobryus in the Caribbean<br />

Trichomycetes are organisms that live within the digestive tracts <strong>of</strong> arthropods.<br />

Historically, they have been considered one class in the Zygomycota, but<br />

molecular studies have determined that the group is not monophyletic. The current<br />

phylogenetic position <strong>of</strong> the orders Amoebidiales and Eccrinales falls within the<br />

protist group Ichthyosporea (also known as Mesomycetozoa). Eccrinales is a morphologically<br />

diverse order with 17 genera, which inhabit a wide range <strong>of</strong> hosts:<br />

crustaceans, beetles and millipedes in varied habitats (marine, freshwater, and terrestrial).<br />

The order is characterized by unbranched, non-septate, multinucleate thalli<br />

and sporangiospores that are formed basipetally from the thallus apex. No species<br />

have been axenically cultured. Our current effort in the Eccrinales is to understand<br />

the diversity and systematics <strong>of</strong> the largest genus, Enterobryus, typically associated<br />

with millipedes. Enterobryus produces two types <strong>of</strong> sporangiospores, a primary<br />

infestation type, which is typically uninucleate, isodiametric and thin-walled, and<br />

a secondary infestation type, which is multinucleate, cylindrical and also thinwalled.<br />

Species delimitation in this genus is extremely difficult with morphology<br />

alone. Intraspecific variation is greater than interspecies variation in some species.<br />

In addition, most species descriptions are based on few collections, which do not<br />

include details <strong>of</strong> character variation. In the Caribbean, most islands have endemic<br />

millipedes and almost all hosts investigated seem to be infected with Enterobryus.<br />

We have been identifying morphological species and including molecular<br />

markers to test species hypothesis. Sequence data from multiple clones <strong>of</strong> 28S<br />

rDNA present significant sequence variation within samples indicating high heterogeneity<br />

in DNA repeats and hence possible relaxation in the concerted evolution<br />

process. New protein markers are needed to generate robust phylogenies.<br />

Capelari, Marina 1 , Fernanda Karstedt 1 , Jadson JS Oliveira 1 , and Nelson<br />

Menolli, Jr. 1,2 1 Instituto de Botânica, Núcleo de Pesquisa em Micologia, Caixa<br />

Postal 68041, 04045-972 São Paulo, SP, Brazil, 2 Instituto Federal de Educação,<br />

Ciência e Tecnologia de São Paulo, Campus São Paulo, CCT / Biologia, Rua<br />

Pedro Vicente, 625, 01109-010 São Paulo, SP, Brazil. A preliminary checklist<br />

<strong>of</strong> Agaricales from Reserva Biológica do Alto da Serra de Paranapiacaba,<br />

Santo André, SP, Brazil<br />

The Atlantic Forest is a long but discontinuous strip <strong>of</strong> tropical rainforest<br />

more than one hundred kilometers wide along the coast, from north to south<br />

Brazil. The “Reserva Biológica do Alto da Serra de Paranapiacaba” located at<br />

23°46´00´´-23°47´10´´S and 46°18´20´´-46°20´40´´W, comprises 336 ha <strong>of</strong> this<br />

forest, with an annual rainfall and temperature <strong>of</strong> 3,300 mm and 14-15°C in the<br />

winter to 21-22°C in the summer, respectively. Altitude varies from 750 m to-891<br />

m and the vegetation comprises high forests, low forests, scrub, dark scrub and<br />

grassland in different successional stages. The fungi were collected randomly<br />

across several trails in the forest, during the period <strong>of</strong> April/2006 to April/2010,<br />

with ca. <strong>of</strong> 367 collections. 140 collections represent 52 species belonging to 18<br />

genera (Calliderma, Collybia, Crepidotus, Cyptotrama, Cystoderma, Gymnopus,<br />

Hypholoma, Inocephalus, Leptonia, Marasmius, Melanotus, Mycena, Nolanea,<br />

Pholiota, Pluteus, Pyrrhoglossum, Ripartitella and Stropharia). The remainder<br />

belong to, at least, 32 genera (Agaricus, Amanita, Camarophyllus, Collybia s.l.,<br />

Conocybe, Coprinus, Crepidotus, Dictyopanus, Favolaschia, Filoboletus, Gerronema,<br />

Hygrocybe, Gloiocephala, Gymnopilus, Gymnopus s.l., Hypholoma, Lepiota,<br />

Leptonia, Leucocoprinus, Marasmiellus, Marasmius, Melanotus, Mycena,<br />

Pholiota, cf. Physocystidium, Pluteus, Psathyrella, Psilocybe, Rimbachia, Trogia,<br />

Tetrapyrgos and Tubaria). New taxa were recorded, some <strong>of</strong> them, already described<br />

(Calliderma fibulatum Karstedt & Capelari, Marasmius cystidioccultus J.<br />

S. Oliveira & Capelari, Marasmius plenicystidiosus J. S. Oliveira & Capelari),<br />

and some are in process <strong>of</strong> publication, especially <strong>of</strong> the genus Marasmius.<br />

Carlson, Alexis L, Alfredo Justo, and David S Hibbett. Clark University. Biology<br />

Department. 950 Main St, 01610, Worcester (MA). Species delimitation in<br />

Trametes (Polyporales, Basidiomycota): a comparison <strong>of</strong> ITS, TEF1, RPB1<br />

and RPB2 phylogenies<br />

Trametes is a cosmopolitan genus <strong>of</strong> white rot polypores, including the<br />

“turkey tail” fungus, T. versicolor. Although Trametes is one <strong>of</strong> the most familiar<br />

genera <strong>of</strong> polypores, its species-level taxonomy is unsettled. The ITS region is the<br />

most commonly used molecular marker for species delimitation in fungi and, despite<br />

some problems and limitations, it generally <strong>of</strong>fers a good level <strong>of</strong> resolution.<br />

During our phylogenetic study <strong>of</strong> the genus Trametes we observed a low level <strong>of</strong><br />

molecular variation in the ITS that resulted in poorly resolved phylogenies and<br />

unclear species boundaries, especially in the Trametes versicolor species group<br />

(T. versicolor, T. ochracea, T. pubescens, T. ectypa). In this study we will evaluate<br />

the performance <strong>of</strong> three protein-coding genes (TEF1, RPB1, RPB2) for<br />

species delimitation and phylogenetic reconstruction in Trametes. Preliminary results<br />

suggest that all three protein-coding genes outperform ITS for separating<br />

species in the T. versicolor complex<br />

Carris, Lori, Kalyn Thomas, Sean McCotter, and Tobin Peever. Department <strong>of</strong><br />

Plant Pathology, Washington State University, Pullman WA 99164-6430. Identification<br />

<strong>of</strong> the anamorphs <strong>of</strong> morels and false morels in Pacific Northwestern<br />

U.S. forests<br />

Morels (Morchella), false morels (Gyromitra) and allied genera are widely<br />

distributed spring macr<strong>of</strong>ungi in the Pacific Northwestern (PNW) US, and recent<br />

studies have supported a western North <strong>America</strong>n origin and diversification<br />

for morels. However, relatively little is known about the biology and ecology <strong>of</strong><br />

this iconic group <strong>of</strong> fungi, and many parts <strong>of</strong> the inland PNW remain under sampled.<br />

As part <strong>of</strong> an ongoing study on the diversity and mating systems <strong>of</strong> black<br />

morels (M. elata group) and allied genera, specimens from 15 morel and false<br />

morel populations were sampled at three locations in northern Idaho and eastern<br />

Washington in May-June 2011. Putative anamorphs were collected at these same<br />

sites in Oct-Dec. Anamorph colonies and ascocarps were photographed, morphologically<br />

characterized, and established in culture. Brown and white anamorph<br />

morphotypes were distinguished based on mycelium color and presence or absence<br />

<strong>of</strong> setae, respectively. Portions <strong>of</strong> ITS rDNA, RPB1, RPB2, and EF1-alpha<br />

were sequenced, and used to confirm anamorph-teleomorph connections for<br />

brown anamorphs with black morel lineages, and white anamorphs with Disciotis<br />

and Gyromitra spp. This is the first report <strong>of</strong> anamorphs for Disciotis and Gyromitra,<br />

and the first report <strong>of</strong> the Morchella anamorph from nature. An anamorphic<br />

state corresponding to the genus Costantinella has been reported in cultured<br />

morels, but nothing is known about the role <strong>of</strong> the anamorph in the morel life<br />

cycle in nature. Identification <strong>of</strong> the widespread occurrence <strong>of</strong> the anamorphs <strong>of</strong><br />

morels and false morels promises to fill major gaps in our understanding <strong>of</strong> the biology<br />

and ecology <strong>of</strong> these fungi.<br />

Carver, Jonathan P and Thomas J Volk. Department <strong>of</strong> Biology, University <strong>of</strong><br />

Wisconsin, La Crosse, WI. Fungal root endophytes in barley, tomato, and eggplant<br />

In recent years it has become apparent that fungal root endophytes are<br />

both ubiquitous in nature and can have significant effects on host plant fitness. In<br />

fact many fungal root endophytes have extraradical hyphae that may allow them<br />

to function in a manner similar to mycorrhizae. Many <strong>of</strong> these extraradical root<br />

endophytes (ERE) are also facultative associates and can be easily cultured in the<br />

lab. The ability <strong>of</strong> these endophytes to be easily cultured, along with their ability<br />

to improve plant growth, means that there is a great potential for using these fungi<br />

to improve agricultural productivity and sustainability. Despite this potential we<br />

still know very little about ERE ecology and diversity and even less about the<br />

EREs naturally present in agricultural systems. In this research project putative<br />

fungal endophytes were isolated from three crop plants grown in garden soil; barley<br />

(Hordeum vulgare), tomato (Solanum lycopersicum), and eggplant (Solanum<br />

melongena). From these plants a total <strong>of</strong> 119 fungal isolates were obtained. Microscopic<br />

traits and culture morphology were used to group isolates into operational<br />

taxonomic units (OTUs) likely equivalent to morphological species or<br />

strains. Twelve <strong>of</strong> the most abundant OTUs were separated for further study and<br />

identified by sequencing the ITS region. These endophytes were subsequently cocultured<br />

with each <strong>of</strong> the three crop plants in axenic growth media under indoor<br />

lights for 30 days. Effect <strong>of</strong> endophyte inoculation on shoot and root biomass was<br />

assessed and endophytes were re-isolated in order to fulfill Koch’s postulates. Additionally,<br />

roots were stained with Trypan Blue and Sudan IV to confirm ERE<br />

habit <strong>of</strong> isolates and to compare percent colonization with effects on plant growth.<br />

The results <strong>of</strong> this study may be useful in improving agriculture by expanding our<br />

current knowledge <strong>of</strong> the specific extraradical root endophytes found in crop<br />

plants and their ecology.<br />

Chen, Ko-Hsuan 1 , Gregory Bonito 1 , Hannah Reynolds 1 , Chris Schadt 2 , and<br />

Rytas Vilgalys 1 . 1 Dept. <strong>of</strong> Biology, Duke University Box 90338, Durham,<br />

27708, 2 Oak Ridge National Laboratory, Department <strong>of</strong> Energy, Oak Ridge TN<br />

37831. From trees to grasses: Arbuscular mycorrhizal fungi community is<br />

shaped by host plant preference<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 13


Arbuscular Mycorrhizal fungi (AMF) are important to terrestrial ecosystems;<br />

however, most <strong>of</strong> AMF community studies focused on AMF associate with<br />

grasses rather than trees. We examined the AMF community on a forest tree, Populus<br />

deltoides, which is associated with ectomycorrhizal fungi as well as AMF. Soil<br />

samples from different localities (North Carolina and Tennessee) were used to inoculate<br />

P. deltoides cuttings. Mycorrhization <strong>of</strong> P. deltoides roots was characterized<br />

using 454 pyrosequencing and cloning <strong>of</strong> the AMF SSU rRNA genes. Subsequently,<br />

fungal communities from P. deltoides roots were transferred to Sorghum bicolor,<br />

a common plant host used for retaining AMF in pot cultures. To detect AMF<br />

community associated with S. bicolor, spores were extracted and identified based on<br />

PCR amplification <strong>of</strong> the SSU rRNA gene using several different primer sets. We<br />

compared data generated by these three approaches (1) P. deltoides roots/ 454 pyrosequencing;<br />

(2) P. deltoides roots/cloning and Sanger sequencing; (3) S. bicolor<br />

roots and rhizosphere /PCR. Each method recovered 35, 10 and 10 OTUs, respectively,<br />

but only 2 common OTUs were detected by all three methods. Combined results<br />

reveal the dominance <strong>of</strong> Glomerales and Paraglomerales in the P. deltoides<br />

AMF community, with few (less than 0.01%) <strong>of</strong> the fungal OTUs detected from the<br />

Diversisporales and none from Archaeosporales. P. deltoides appears to be preferably<br />

mycorrhized by Glomerales and Paraglomerales. Several Rhizophagus OTUs<br />

were only detected by S. bicolor/PCR method, suggesting Rhizophagus fungi might<br />

exist in Populus roots in very low numbers, but are favored by S. bicolor once this<br />

host is available. These results suggest that Sorghum may not always be an optimal<br />

alternate host for culturing AMF fungi, particularly those that preferentially associate<br />

with Populus. Currently we are investigating the transferability <strong>of</strong> AMF communities<br />

from S. sorghum back to P. deltoides.<br />

Choi, Jong-In 1 , Jeon Dae-Hoon 1 , Tai- Moon Ha 1 , Young-Cheul Ju 1 and Geon-<br />

Sik Seo 2 . 1 Mushroom Research Institute, Gyeonggido Agiricultural Research &<br />

Extension Services, 430-8 Sam-Ri Goniliam-Eup Gwangju-Si, Gyeonggi-Do<br />

464-873, Republic <strong>of</strong> Korea. 2 Korea National College <strong>of</strong> Agriculture and Fisheries,<br />

11-1 Donghwa-Ri Bongdam-Eup Hwaseong-Si Kyeonggi-Do 445-760,<br />

Republic <strong>of</strong> Korea. Characteristics <strong>of</strong> a new mid-high temperature adaptable<br />

oyster mushroom variety ‘Gonji-5ho’ for bag culture<br />

‘Gonji-5ho,’ a new variety <strong>of</strong> oyster mushroom that is suitable for production<br />

in bag culture, was bred by matings between monokaryons isolated from<br />

chiak-3ho and Suhan-1ho. The major characteristics <strong>of</strong> the fruit body included a<br />

thick and gray pileus and thick, long and s<strong>of</strong>t stipes. Tissue elasticity and cohesivness<br />

was superior to Suhan-1ho. The optimum temperature for mycelial<br />

growth was around 26 to 29C and that for pinheading and growth <strong>of</strong> fruitbodies<br />

was around 18 to 20C. In the bag culture, it 18 days <strong>of</strong> incubation was required<br />

and 3 days for the formation <strong>of</strong> primordia. The fruit bodies grew to maturity and<br />

were uniform in size, yielding 221.4g/1kg bag.<br />

Cifuentes, Joaquín B and Rosalva A Vázquez-Estup. Biología, Facultad de<br />

Ciencias, UNAM, México DF. Cystolepiota and Echinoderma: tropical spp. in<br />

México<br />

Cystolepiota as currently defined (Vellinga, 2007) contains between 10<br />

and 55 taxa (Kirk et al., 2008; Vellinga, 2010). Echinoderma is considered either<br />

a synomyn <strong>of</strong> Lepiota (Kirk et al., op. cit., Vellinga, 2010) or a separate genus<br />

with up to 14 taxa (Bon, 1992; Knudsen et Versterholt, 2008). Phylogenetic<br />

analysis suggested a close relationship between these genera (Vellinga, 2003).<br />

Only the following species have been reported from Mexico (under different genera):<br />

C. acutesquamosa, C. cygnea and L. hemisclera (Herrera and Guzmán,<br />

1972; Bandala et al., 1988; Cifuentes, 2008). Our ongoing revision <strong>of</strong> lepiotaceous<br />

fungi from Mexico has revealed some interesting tropical taxa in these genera.<br />

Here we describe and illustrate two taxa in Cystolepiota (one C. aff. petasiformis<br />

and a putative new taxon) and two putative new taxa in Echinoderma.<br />

Corradi, Nicolas and Stefan Amyotte. University <strong>of</strong> Ottawa, Ottawa, Ontario,<br />

Canada. Comparative transcriptomics and the arbuscular mycorrhizal fungi<br />

Arbuscular mycorrhizal fungi (AMF) are notorious for benefiting most<br />

land plants by establishing a widespread association with their roots; the mycorrhizal<br />

symbiosis. Despite their tremendous importance for terrestrial plants worldwide,<br />

however, the content and nature <strong>of</strong> their genomes has long remained elusive.<br />

Here, we report the acquisition <strong>of</strong> large-scale sequence data from the<br />

transcriptome <strong>of</strong> two different strains <strong>of</strong> AMF in the genus Rhizophagus (previously<br />

known as Glomus) using Illumina sequencing. A total <strong>of</strong> 39,000 contigs<br />

from these strains were assembled, annotated and compared against sequence data<br />

(over 40,000 contigs) recently acquired by others from the model AMF Rhizophagus<br />

(Glomus) intraradices. Overall, these high-throughput approaches resulted<br />

in the first large-scale comparison <strong>of</strong> biochemical processes occurring among<br />

closely related AMF strains; providing first-hand and long-awaited information<br />

about the diversity, evolution, origin and plasticity <strong>of</strong> the AMF proteome. In particular,<br />

our study revealed a number <strong>of</strong> evolutionary novelties and molecular<br />

mechanisms that were previously unknown to occur in the genome <strong>of</strong> these ecologically<br />

critical organisms.<br />

14 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Crandall, Sharifa G and Gregory S Gilbert. University <strong>of</strong> California, Santa Cruz,<br />

1156 High Street, Santa Cruz, CA 95064. Airborne fungal spore dispersal and<br />

trait diversity in coastal mixed-evergreen forests in California<br />

Although some studies characterize fungal spore dispersal dynamics in<br />

tropical forests, spore dynamics in temperate forests is still largely unknown. Airborne<br />

spore abundance in tropical forests is stratified: there is a higher density <strong>of</strong><br />

spores in the still, dark, and moist understory compared to the windy, light, and<br />

dry conditions <strong>of</strong> the forest canopy. This spatial stratification suggests that spore<br />

dispersal is influenced by abiotic factors and that species may be dispersal limited.<br />

Moreover, once spores are discharged, specific spore traits such as size or ornamentation<br />

may restrict or facilitate spore movement. This study examines the<br />

spatial dynamics <strong>of</strong> airborne spores in a temperate coastal mixed-evergreen forest<br />

in California. We ask: 1) is there is a difference in spore density under different<br />

site conditions (open vs. closed canopy)? 2) do fungal spore traits differ among<br />

sites and vertically from the forest floor into the canopy? 3) which spore traits are<br />

phylogenetically conserved across taxa? Spores from different fungal species<br />

were trapped in closed versus open canopy forest sites at the University <strong>of</strong> California,<br />

Santa Cruz Upper Campus Reserve during the wet season in 2012. Spore<br />

density and traits are characterized by imaging <strong>of</strong> spores under a light microscope<br />

and analyzed using the s<strong>of</strong>tware analysis tool imageJ®. Traits measured include<br />

spore diameter, size, outer ornamentation, cell wall thickness, and presence/absence<br />

<strong>of</strong> melanin. Spore samples will be identified using next generation sequencing<br />

(NGS) methods. We report on the spatial distribution <strong>of</strong> airborne spores<br />

and suggest phylogenetically conserved traits across fungal communities.<br />

Cripps, Cathy L 1 , Erin Lonergan 1 , and Cyndi M Smith 2 . 1 Plant Sciences and<br />

Plant Pathology Department, Montana State University Bozeman MT 59717,<br />

USA, 2 Parks Canada, Waterton Lakes National Park, P.0. 200, Waterton Park,<br />

AB TOK 2MO, Canada. Linking fungal biology to restoration: Use <strong>of</strong> hostspecific<br />

ectomycorrhizal fungi to restore whitebark pine, a ‘threatened’<br />

species in western North <strong>America</strong><br />

North <strong>America</strong>’s only stone pine (Pinus albicaulis) is under consideration<br />

for ‘endangered species’ status in the USA and is assessed as ‘endangered’ in<br />

Canada awaiting legal listing. This five-needle pine forms extensive forests <strong>of</strong><br />

magnificent old growth or gnarled krummholz at treeline in western North <strong>America</strong>.<br />

It has a unique ecology that requires bird dispersal <strong>of</strong> seeds; squirrels compete<br />

for seeds and grizzly bears raid squirrel caches in a complicated food web that includes<br />

ectomycorrhizal fungi (EMF). The invasive white pine blister rust and<br />

mountain pine beetle have reduced forests 12-70% in parts <strong>of</strong> their original range.<br />

Large restoration efforts are underway as summarized in the 2012 ‘Range-Wide<br />

Restoration Strategy for Whitebark Pine’ that promotes plantings <strong>of</strong> rust-resistant<br />

nursery seedlings. However survival rates <strong>of</strong> previously planted seedlings are low.<br />

We are investigating the use <strong>of</strong> native EMF as a tool to enhance survival for various<br />

projects including “Working together to Restore Terrestrial Ecosystems” for<br />

Parks Canada. Our surveys <strong>of</strong> EMF with this pine have revealed limited EMF diversity<br />

and an important set <strong>of</strong> suilloids. Some <strong>of</strong> the suilloids also occur with<br />

stone pines in Europe and Asia; others appear limited to whitebark pine. Host-specific<br />

fungi could confer an advantage not available to competitive trees. However<br />

host-specific fungi are also at risk and could decline with the tree species making<br />

restoration more difficult. Screening <strong>of</strong> 26 native EMF showed Suillus species<br />

related to those used to inoculate cembrean pine in Europe to be effective colonizers<br />

in the nursery; Rhizopogon lagged in colonization. We are currently developing<br />

inoculation methods for effective colonization <strong>of</strong> this tree species. Large<br />

out-plantings <strong>of</strong> inoculated seedlings are being monitored in Waterton Lakes National<br />

Park to determine if survival is enhanced. Restoration recommendations<br />

specific to whitebark pine will be presented with preservation strategies for hostspecific<br />

EMF.<br />

Cui, Yunluan and Nicholas P Money. Miami University 316 Pearson Hall Oxford<br />

Ohio 45056. Actin dynamics in constricting ring traps formed by nematophagous<br />

fungi<br />

The micr<strong>of</strong>ilament network <strong>of</strong> F-actin plays an essential role in fungal<br />

morphogenesis by coordinating cell wall synthesis, cytoplasmic migration, and<br />

organelle positioning. We are investigating the connection between the behavior<br />

<strong>of</strong> the actin cytoskeleton and trap development in nematophagous fungi that form<br />

constricting ring traps. We have constructed a fluorescence-tagged actin-binding<br />

protein (Lifeact) driven by the ToxA promoter through recombinant PCR. The<br />

vector pBChygro harboring an ampicillin resistant gene and hygromycin B phosphotransferase<br />

gene was used as a backbone to accept the chimeric gene into multiple<br />

clone sites through enzyme digestion and linkage. Restriction enzyme mediated<br />

integration (REMI) and the CaCl2-polyethylene glycol methods were used<br />

to transform fungal protoplasts with the Lifeact-expressing plasmid. This efficient<br />

transformation system facilitated the random insertion <strong>of</strong> the HindIII linearized<br />

plasmid into the fungal genome. Transformed protoplasts were imaged using flu-<br />

Continued on following page


orescence microscopy. Treatment with restriction enzymes had no effect on protoplast<br />

survival and hyphae generated from protoplasts expressed the fluorescent<br />

protein tag. This indicates that heterogeneous Lifeact protein is expressed under<br />

the ToxA promoter, and that integration <strong>of</strong> the plasmid into the fungal genome was<br />

successful. This is the first report <strong>of</strong> the successful transformation <strong>of</strong> a fungus that<br />

produces ring traps using heterogeneous Lifeact by REMI. Data obtained from<br />

these experiments will be helpful in the future investigation <strong>of</strong> functional genes<br />

contributing to trap development.<br />

Davey, Marie L 1,2 , Håvard Kauserud 2 , and Mikael Ohlson 1 . 1 Department <strong>of</strong><br />

Ecology and Natural Resource Management, Norwegian University <strong>of</strong> Life Sciences,<br />

N-1432 Ås, Norway, 2 Molecular Evolution Research Group, Department<br />

<strong>of</strong> Biology, University <strong>of</strong> Oslo, N-0316 Oslo, Norway. Amplicon pyrosequencing<br />

reveals shifts in the fungal communities associated with boreal<br />

bryophytes along an elevation and vegetation-type gradient<br />

Bryophytes are a dominant vegetation component <strong>of</strong> the boreal and alpine<br />

ecosystems and are thought to host a high diversity <strong>of</strong> fungi that play important<br />

ecological roles in nutrient cycling. However, relatively little is known about fungal<br />

community size and composition dynamics in these systems. Four bryophyte<br />

hosts (Dicranum scoparium, Hylocomium splendens, Pleurozium schreberi, and<br />

Polytrichum formosum) were collected along elevational transects spanning three<br />

distinct vegetation zones: low alpine meadow, sub-alpine birch forest, and montane<br />

conifer forest. 454 amplicon pyrosequencing <strong>of</strong> the ITS2 region <strong>of</strong> rDNA<br />

was used to characterize the fungal communities associated with these mosses and<br />

to investigate community dynamics in composition and diversity along the gradient.<br />

The biochemical marker compound ergosterol was quantified and used as a<br />

proxy for fungal biomass to allow for estimations <strong>of</strong> the size <strong>of</strong> these fungal communities.<br />

The overall fungal biomass associated with the moss communities varied<br />

between both the host species and the photosynthetic and senescent tissues <strong>of</strong><br />

the moss but remained constant across the different vegetation types. Although<br />

community size (fungal biomass) remained constant across the gradient, distinct<br />

shifts in community composition occurred between the three vegetation types,<br />

with a transition occurring between the community associated with the hosts in<br />

the subalpine birch vegetation type and those in the montane conifer forest type.<br />

The four moss hosts displayed varying degrees <strong>of</strong> patchiness and spatial auto-correlation<br />

in their associated fungal communities.<br />

Davidson, Bill E and Marcelo D Serpe. Department <strong>of</strong> Biological Sciences, 1910<br />

University Drive, Boise State University, Boise, ID 83725-1515. Improvement<br />

in colonization and seedling survival <strong>of</strong> Wyoming big sagebrush following inoculation<br />

with native arbuscular mycorrhizal fungi<br />

Inoculation <strong>of</strong> seedlings with arbuscular mycorrhizal fungi (AMF) is a<br />

common practice aimed at improving seedling establishment. The success <strong>of</strong> this<br />

practice largely depends on the ability <strong>of</strong> the inoculum to colonize the growing<br />

root system after transplanting. These events were investigated in Wyoming big<br />

sagebrush (Artemisia tridentata ssp. wyomingensis) seedlings inoculated with native<br />

AMF. Seedlings were first grown in a greenhouse in 50 mL containers containing<br />

sterilized pot cultures (control seedlings) or pot cultures having a mixture<br />

<strong>of</strong> seven mycorrhizal phylotypes (inoculated seedlings). In early spring, threemonth<br />

old seedlings were transplanted to 24 L pots filled with soil collected from<br />

a sagebrush habitat and grown under natural climatic conditions. At the time <strong>of</strong><br />

transplanting the percent colonization was 0.3 and 57% for control and inoculated<br />

seedlings, respectively. Two and a half and five months after transplanting, the<br />

percent colonization remained higher in inoculated than in control seedlings with<br />

values <strong>of</strong> about 48 and 20%, respectively. These differences in colonization were<br />

associated with differences in seedling survival, which was 24% higher in inoculated<br />

than control seedlings. In contrast, biomass per plant and the shoot over root<br />

ratio were similar for both treatments. Overall, our results indicate that the inoculum<br />

contributed to the colonization <strong>of</strong> the roots that developed after transplanting<br />

resulting in higher levels <strong>of</strong> colonization than those naturally occurring in the soil.<br />

The increased colonization by mycorrhizal fungi may have led to an increase in<br />

seedling survival.<br />

Davis, William J, Martha J Powell, Peter M Letcher, and Jonathan Antonetti. Department<br />

<strong>of</strong> Biological Sciences, The University <strong>of</strong> Alabama, Tuscaloosa, AL,<br />

35487. Exploring sporangial surface features to characterize members <strong>of</strong> the<br />

Chytriomycetaceae (Chytridiales)<br />

Understanding species boundaries is critical to understanding biodiversity.<br />

In the past plasticity <strong>of</strong> the chytrid thallus led to the consolidation <strong>of</strong> morphologically<br />

variable isolates into a single species. Recent molecular phylogenetic<br />

analyses have <strong>of</strong>ten revealed great genetic diversity among isolates and validated<br />

the separation <strong>of</strong> these isolates into distinct species. It has also been revealed that<br />

morphologically similar isolates can show great genetic diversity. This is true <strong>of</strong><br />

Chytriomyces hyalinus, one <strong>of</strong> the most commonly collected chytrids. Isolates<br />

identified as C. hyalinus based on thallus morphology and habitat form four wellsupported<br />

clades in molecular analyses. These results suggest the presence <strong>of</strong><br />

cryptic species. Characters that distinguish these potential cryptic species need to<br />

be identified in order to delineate and describe them as new species. Examination<br />

<strong>of</strong> sporangial cell surface features with scanning electron microscopy (SEM) has<br />

revealed that many <strong>of</strong> these isolates <strong>of</strong> C. hyalinus have ridge-like surfaces, which<br />

are not readily detected with light microcopy and were not observed in the original<br />

description <strong>of</strong> C. hyalinus. The purpose <strong>of</strong> this study is to explore whether sporangial<br />

surface features can be used to distinguish members in different sub-clades<br />

<strong>of</strong> the C. hyalinus complex. Thus, representatives <strong>of</strong> each <strong>of</strong> the sub-clades <strong>of</strong> the<br />

C. hyalinus complex, as well as sister clades, were examined with SEM.<br />

Davoodian, Naveed. New York Botanical Garden, Institute <strong>of</strong> Systematic<br />

Botany, 2900 Southern Blvd, Bronx, NY 10458. Toward a monograph <strong>of</strong> Gyroporus<br />

The genus Gyroporus is a group <strong>of</strong> poroid ectomycorrhizal mushrooms in<br />

the Boletales with representatives documented from every major continent except<br />

Antarctica. They are strongly implicated as symbionts with Pinaceae, Fagaceae,<br />

Myrtaceae, and Casuarinaceae and there is evidence for associations with Betulaceae,<br />

Salicaceae, Euphorbiaceae, and Fabaceae as well. Though the genus is<br />

fairly easily distinguished (by a yellow spore print, broadly circumferentially<br />

arranged stipe hyphae, and clamp connections), and some members have been<br />

known since before Fries, the group has never been monographed. Research is<br />

currently underway on a contribution toward a monograph <strong>of</strong> Gyroporus. A preliminary<br />

overview <strong>of</strong> the research is presented, with emphasis on problems in the<br />

study <strong>of</strong> this group. Several major nomenclatural problems exist within the group.<br />

Also, preliminary phylogenetic studies by other investigators have yielded equivocal<br />

results; this is further compounded by the problem <strong>of</strong> morphologically similar<br />

taxa exhibiting disjunct, multi-continental distributions. As such, some <strong>of</strong> the<br />

well-accepted species in the genus are suspected to actually be species complexes.<br />

A global approach with broader sampling has been adopted to address these<br />

issues, and recent fieldwork in Australia has generated important collections and<br />

likely new species.<br />

de Beer, Z Wilhelm 1 , Keith A Seifert 2 , and Michael J Wingfield 1 . 1 Department<br />

<strong>of</strong> Microbiology and Plant Pathology, FABI, University <strong>of</strong> Pretoria, Pretoria,<br />

South Africa, 2 Biodiversity (Mycology and Botany), Agriculture and Agri-Food<br />

Canada, Ottawa, Ontario, K1A 0C6, Canada. The impacts <strong>of</strong> 1F:1N on the<br />

Ophiostomatales (Sordariomycetes, Ascomycota)<br />

The Ophiostomatales includes mainly bark beetle associated fungi, the<br />

majority <strong>of</strong> which are responsible for blue stain <strong>of</strong> freshly exposed timber, while<br />

a few species are tree-killing pathogens. Under the dual nomenclature system,<br />

four teleomorph genera were recognized in the Ophiostomatales: Ophiostoma,<br />

Ceratocystiopsis, Grosmannia, and Fragosphaeria. Anamorphs in the order were<br />

treated in recent years primarily in Sporothrix, Leptographium, Pesotum, Raffaelea<br />

and Hyalorhinocladiella, while seven additional anamorph genera were treated<br />

as synonyms <strong>of</strong> these five genera. The adoption <strong>of</strong> the one fungus one name<br />

(1F:1N) principles in the ICN has ‘released’ the unused genus names for application<br />

to newly emerging lineages in the Ophiostomatales, which include species<br />

with known teleomorphs. The aims <strong>of</strong> this study were to consider the impact <strong>of</strong><br />

the application <strong>of</strong> the 1F:1N principles on the Ophiostomatales, to redefine genera,<br />

to reconsider synonymies and to resurrect generic names where applicable.<br />

To achieve this, we re-assessed all published rDNA sequences for the Ophiostomatales,<br />

selected the most reliable sequences representing 269 taxa, and conducted<br />

phylogenetic analyses using these data. The resulting phylogenies confirmed<br />

the monophyly <strong>of</strong> Ceratocystiopsis and Fragosphaeria, but also showed<br />

that Ophiostoma, Grosmannia, and Raffaelea are polyphyletic. The S. schenckii-<br />

O. stenoceras- and P. fragrans species complexes, previously treated in Ophiostoma,<br />

emerged as distinct genera, namely Sporothrix and Graphilbum. We thus<br />

redefined both these genera to include species with teleomorphs previously assigned<br />

to Ophiostoma. However, presently available data are insufficient to clarify<br />

the generic status <strong>of</strong> several other lineages currently treated in Ophiostoma s.l.,<br />

including the status <strong>of</strong> Leptographium versus Grosmannia, or that <strong>of</strong> the R. lauricola<br />

complex. Because it is presently not possible to propose definitive generic<br />

names for some clades, the relevant species are left in their current classification<br />

in order to minimize nomenclatural changes, pending the selection <strong>of</strong> acceptable<br />

generic names.<br />

Dentinger, Bryn TM 1,2 , D Jean Lodge 3 , A Martyn Ainsworth 1 , Paul F Cannon<br />

1 , and Gareth W Griffith 2 . 1 Jodrell Laboratory, Royal Botanic Gardens, Kew,<br />

Richmond, Surrey TW9 3DS, UK, 2 Institute <strong>of</strong> Biological, Environmental and<br />

Rural Sciences, Prifysgol Aberystwyth, Aberystwyth, Ceredigion Wales SY23<br />

3DD, UK, 3 Center for Forest Mycology Research, US Department <strong>of</strong> Agriculture,<br />

Forest Service, Northern Research Station, P.O. Box 1377, Luquillo, Puerto<br />

Rico 00773. Systematics and DNA barcoding <strong>of</strong> waxcap fungi<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 15


Fungi are crucial components <strong>of</strong> all ecosystems and their conservation requires<br />

strategies specifically tailored to them. However, a major stumbling block<br />

in fungal conservation is our lack <strong>of</strong> knowledge <strong>of</strong> their true diversity. In the UK,<br />

waxcap fungi (Hygrocybe s.l. and Cuphophyllus) have been the subject <strong>of</strong> longterm<br />

monitoring, they have excited the attention <strong>of</strong> non-specialists, their presence<br />

has led to designation <strong>of</strong> a few Special Sites <strong>of</strong> Scientific Interest in the UK, and<br />

they are included in the sites analyzed in the Important Fungus Areas survey.<br />

Moreover, all four national statutory conservation bodies have funded survey and<br />

monitoring <strong>of</strong> these fungi. However, over the last decades, application <strong>of</strong> molecular<br />

markers to species diagnosis in fungi has called into question the reliability<br />

<strong>of</strong> a strictly morphology-based approach. A key requirement for many fungi <strong>of</strong><br />

conservation concern in the UK is the need for molecular diagnostic tools to assist<br />

in species definition (including recognition <strong>of</strong> cryptic taxa). For this project,<br />

we focused on generating DNA barcodes from waxcap fungi to assist in identification<br />

<strong>of</strong> cryptic species that may need to be considered for conservation management.<br />

New citizen-led survey work has contributed over 500 new collections<br />

across Great Britain. In addition, sequencing information was harvested from over<br />

650 fungarium specimens representing most <strong>of</strong> the known species in the UK.<br />

Over 200 new ITS barcode sequences were analyzed in combination with all <strong>of</strong><br />

the Hygrocybe ITS sequences available on GenBank plus unpublished reference<br />

sequences, revealing a number <strong>of</strong> cryptic species and species with identification<br />

problems in the UK and USA. A reevaluation <strong>of</strong> morphological characters correlated<br />

with distinct ITS clades will inform future rapid, field-based surveys and<br />

monitoring efforts. In addition, these sequences have helped in revising infrageneric<br />

names in Hygrophoraceae - the first full revision since Hesler and Smith<br />

(19<strong>63</strong>).<br />

Desai, Nikhilesh S. Chicago Botanic Garden, Plant Science Center, 1000 Lake<br />

Cook Rd, Glencoe, IL 60022. Mycorrhizal community composition <strong>of</strong> Quercus<br />

oleoides as a function <strong>of</strong> stand maturity in a regenerating dry tropical<br />

forest<br />

Species rich dry tropical forests (DTFs) are the most endangered ecosystems<br />

in the tropical biome, threatened by many anthropogenic activities. In Costa<br />

Rica, deforestation <strong>of</strong> the DTF has occurred for centuries; however, conservation<br />

efforts over the past 30 years have restored these forests to 47.9% <strong>of</strong> their original<br />

extent. To better understand the role <strong>of</strong> mycorrhizal fungi in the regeneration<br />

<strong>of</strong> DTFs, this project studies the mycorrhizal communities associated with a<br />

younger (10 years old) and older (25 years old) stand <strong>of</strong> Quercus oleoides (tropical<br />

live oak). The study site is Sector Santa Rosa <strong>of</strong> the Área de Conservación<br />

Guanacaste (ACG) in northwestern Costa Rica, where Q. oleoides were once regionally<br />

dominant. This project is intended to evaluate: 1) the ectomycorrhizal<br />

(EM) diversity in older and younger stands <strong>of</strong> oak and the overlap in community<br />

composition between stands; 2) whether older stands contain more rare EM<br />

species than younger stands; and 3) how the balance between arbuscular mycorrhizal<br />

(AM) and EM abundance reflects stand age. Data collection occurred over<br />

two weeks in July 2011. Using two 20-m x 50-m plots from young and old stands,<br />

soil cores were collected and the associated EM root tips were pooled for analysis<br />

<strong>of</strong> species composition using molecular methods. Sporocarps were collected<br />

for DNA analysis <strong>of</strong> fresh fungal tissue. DNA sequences were used to identify<br />

EM fungi using BLAST searches against GenBank and UNITE databases. Oak<br />

roots without mantles were analyzed for AM presence by staining with Trypan<br />

blue. More EM root tip colonization was observed in soil cores from older stands.<br />

Preliminary RFLP results suggest that this reflects a greater EM richness in these<br />

communities. Further results and discussion will be presented in the poster.<br />

Li, De-Wei 1 and Guihua Zhao 2 . 1 The Connecticut Agricultural Experiment Station,<br />

Valley Laboratory, 153 Cook Hill Road, Windsor, CT 06095, 2 Center <strong>of</strong><br />

Biotechnology R&D, Jiangsu Polytechnic College <strong>of</strong> Agriculture and Forestry, 3<br />

Jushu Road, Jurong, Jiangsu 212400, China. Two new hyphomycetes, Strelitziana<br />

windsorensis sp. nov. and Zeloasperisporium toonae sp. nov., from<br />

Toona sinensis in Connecticut<br />

Two new hyphomycetes, Strelitziana windsorensis and Zeloasperisporium<br />

toonae, that were collected from Toona sinensis in Windsor, Connecticut were<br />

found to be new to science according to the morphological characters and analysis<br />

<strong>of</strong> ITS sequences. Strelitziana windsorensis is described and illustrated with<br />

filiform conidia, smooth, <strong>63</strong>-103 _ 2.0-2.5 _m and conidiophores reduced to conidiogenous<br />

cells and Zeloasperisporium toonae with conidia, smooth, 12.3-17 µ<br />

3.5-4.4 µm, 0-2 (1) septate. Keys to the species <strong>of</strong> Strelitziana and Zeloasperisporium<br />

are provided.<br />

van Diepen, Linda TA 1 , Serita D Frey 1 , and W Kelley Thomas 2 . 1 Department<br />

<strong>of</strong> Natural Resources and the Environment, University <strong>of</strong> New Hampshire,<br />

Durham, NH 03824, 2 Hubbard Center for Genome Studies, University <strong>of</strong> New<br />

Hampshire, Durham, NH 03824. Soil metatranscriptomics reveals changes in<br />

expression <strong>of</strong> transcripts encoding lignocellulolytic enzymes in the forest<br />

floor <strong>of</strong> a temperate forest under increased nitrogen deposition<br />

16 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Fungi are ubiquitous in terrestrial ecosystems and play an important role<br />

in biogeochemical cycling because <strong>of</strong> their function as litter decomposers. It has<br />

been demonstrated that increased nitrogen (N) deposition decreases fungal biomass<br />

and changes the relative abundance <strong>of</strong> particular groups and species. In addition<br />

increased N can slow litter decomposition and reduce lignolytic enzyme activity.<br />

To understand the functioning <strong>of</strong> fungi under increased N addition in more<br />

detail, we performed a soil metatranscriptomic analysis on forest floor samples<br />

from a chronic N-addition experiment at Harvard Forest (MA, USA). We extracted<br />

total RNA, followed by reverse transcription <strong>of</strong> poly-A tailed mRNA and<br />

Illumina paired-end sequencing <strong>of</strong> the cDNA. In the forest floor <strong>of</strong> control and N<br />

amended plots, a total <strong>of</strong> ~460,000 genes were expressed at the time <strong>of</strong> sampling,<br />

<strong>of</strong> which 172,000 and 95,000 were uniquely expressed in the N addition and control<br />

plots, respectively. More than 20,000 genes (~5% <strong>of</strong> all expressed genes)<br />

were either significantly up-regulated or down-regulated in the control versus the<br />

N-amended plots (P5, Reads Per Kilobase <strong>of</strong> exon per Million<br />

mapped sequence reads) in either the control or N-amended plots. Most <strong>of</strong> the differences<br />

in expression were found within primary metabolic and cellular processes.<br />

More specifically, several genes closely related to fungal glycoside hydrolases<br />

and laccases, involved in lignocellulose degradation, were found to be down-regulated<br />

in the N-amended plots compared to the control plots. Within our study,<br />

metatranscriptomics proved to be a useful technique to study the activity <strong>of</strong> specific<br />

genes <strong>of</strong> interest, while at the same time revealing shifts in the activity <strong>of</strong><br />

genes that would be overlooked in gene-specific qPCR analysis.<br />

Dillon, Husbands R 1 and Henkel W Terry 2 . 1 Department <strong>of</strong> Agriculture, Faculty<br />

<strong>of</strong> Agriculture and Forestry, University <strong>of</strong> Guyana, Georgetown, Guyana, 2 Department<br />

<strong>of</strong> Biological Sciences, Humboldt State University, Arcata, CA 95521.<br />

Synopsis <strong>of</strong> Boletaceae from the central Guiana Shield<br />

<strong>Mycological</strong> exploration in the remote tropical forests <strong>of</strong> Guyana, in the<br />

central Guiana Shield phytogeographic region <strong>of</strong> northeastern South <strong>America</strong>, has<br />

uncovered upwards <strong>of</strong> 30 species <strong>of</strong> the pore mushroom family Boletaceae sensu<br />

lato. Boletes have heret<strong>of</strong>ore been poorly documented from the South <strong>America</strong>n<br />

tropics. These species are distributed across eight sensu lato genera including Austroboletus,<br />

Boletellus, Chalciporus, Fistulinella, Phylloporus, Pulveroboletus,<br />

Tylopilus, and Xerocomus. Eighteen <strong>of</strong> these species, most <strong>of</strong> which were new to<br />

science, have been formally described. Molecular analysis <strong>of</strong> roots has confirmed<br />

that at least sixteen <strong>of</strong> these bolete species are ECM associates <strong>of</strong> the Fabaceae<br />

hosts Dicymbe spp., Aldina insignis, the endemic dipterocarp Pakaraimaea dipterocarpacea,<br />

or Coccoloba spp. (Polygonaceae). We will discuss current work on<br />

new species <strong>of</strong> Xerocomus from this bolete-rich region.<br />

Duong, Tuan A 1 , Z Wilhelm de Beer 2 , Brenda D Wingfield 1 , and Michael J<br />

Wingfield 1 . 1 Department <strong>of</strong> Genetics, Forestry and Agricultural Biotechnology<br />

Institute, University <strong>of</strong> Pretoria, Pretoria 0002, South Africa, 2 Department <strong>of</strong> Microbiology<br />

and Plant Pathology, Forestry and Agricultural Biotechnology Institute,<br />

University <strong>of</strong> Pretoria, Pretoria 0002, South Africa. Cloning, characterization<br />

and population analysis <strong>of</strong> the mating-type genes from Leptographium<br />

procerum and L. pr<strong>of</strong>anum<br />

Leptographium procerum and its sibling species, L. pr<strong>of</strong>anum, are ascomycetes<br />

associated with root-infesting beetles <strong>of</strong> respectively pines and hardwood<br />

trees. Both species are native to North <strong>America</strong>, although L. procerum has<br />

been introduced into Europe, New Zealand, and South Africa. The latest introduction<br />

<strong>of</strong> L. procerum was into China in association with the red turpentine beetle<br />

(Dendroctonus valens), where the fungus apparently contributes to killing <strong>of</strong><br />

millions <strong>of</strong> native Chinese pine trees. As for many other Leptographium species,<br />

sexual states have never been observed in L. procerum and L. pr<strong>of</strong>anum. The objectives<br />

<strong>of</strong> this study were to clone and characterize the mating type loci <strong>of</strong> these<br />

fungi, and to develop markers that can be used to characterize the mating-type <strong>of</strong><br />

individual isolates. To achieve this, we amplified a partial sequence <strong>of</strong> MAT1-2-<br />

1 using degenerate primers targeting the high mobility group (HMG) box. A complete<br />

MAT1-2 idiomorph <strong>of</strong> L. procerum was subsequently obtained by screening<br />

a genomic library using the HMG sequence as a probe. Long range PCRs were<br />

used to amplify the complete MAT1-1 idiomorph <strong>of</strong> L. procerum and both the<br />

MAT1-1 and MAT1-2 idiomorphs <strong>of</strong> L. pr<strong>of</strong>anum. Characterization <strong>of</strong> the MAT<br />

idiomorphs suggested that the MAT genes are fully functional and that individuals<br />

<strong>of</strong> both these species are self-sterile in nature with a heterothallic mating system.<br />

Mating type markers were developed and tested on a population <strong>of</strong> L. procerum<br />

isolates from the USA, the assumed center <strong>of</strong> origin for this species. The<br />

results suggested that cryptic sex is occurring or has recently taken place within<br />

this population. The information regarding the MAT genes and mating type markers<br />

developed in this study will now be used in studies considering the population<br />

genetics and origin <strong>of</strong> these fungi.<br />

Continued on following page


Duong, Tuan A 1 , Z Wilhelm de Beer 2 , Brenda D Wingfield 1 , and Michael J<br />

Wingfield 1 . 1 Department <strong>of</strong> Genetics, Forestry and Agricultural Biotechnology<br />

Institute, University <strong>of</strong> Pretoria, Pretoria 0002, South Africa., 2 Department <strong>of</strong> Microbiology<br />

and Plant Pathology, Forestry and Agricultural Biotechnology Institute,<br />

University <strong>of</strong> Pretoria, Pretoria 0002, South Africa. Characterization <strong>of</strong><br />

MAT genes reveals unexpected patterns <strong>of</strong> sexual compatibility in Leptographium<br />

and Grosmannia<br />

Species <strong>of</strong> Leptographium sensu lato (Ophiostomatales, Ascomycetes)<br />

are sap-stain fungi vectored by bark beetles (Coleoptera, Scolytinae) and some<br />

cause or are associated with root diseases in conifers. Sexual states have been reported<br />

for more than 30 species, and following the dual nomenclature system,<br />

these have been treated in the teleomorph genus Grosmannia. No sexual state is<br />

known for at least 47 additional species and these reside in the anamorph genus<br />

Leptographium. The discovery <strong>of</strong> sexual states for species <strong>of</strong> Leptographium relies<br />

mainly on the presence <strong>of</strong> fruiting bodies on host tissue at the time <strong>of</strong> isolation<br />

and/or intensive laboratory mating studies, which are <strong>of</strong>ten undirected and<br />

labor intensive with a low success rate. Very little is known regarding the mating<br />

systems in Leptographium and Grosmannia and the aim <strong>of</strong> this study was to investigate<br />

the sexual compatibility <strong>of</strong> species in these two genera by characterization<br />

<strong>of</strong> the mating type loci. To achieve this objective, we cloned and identified<br />

mating type genes in six selected species by screening genomic libraries in combination<br />

with long range PCR. Based on these sequences, primers were designed<br />

to amplify MAT genes in other Grosmannia and Leptographium species. The<br />

primers successfully amplified portions <strong>of</strong> the mating type genes <strong>of</strong> more than 30<br />

species and confirmed their thallism, in many species for the first time. Populations<br />

<strong>of</strong> several species were also investigated to obtain evidence <strong>of</strong> sexual reproduction<br />

by comparing the mating type ratio in available isolates. The results suggested<br />

that sexual reproduction probably occurs in nature for several <strong>of</strong> these<br />

species, even though their sexual states have never been seen. The mating type<br />

genes and primers developed in this study represent important tools to further<br />

study the biology and population diversity in the Ophiostomatales.<br />

Eberlein, Chris, Angela M Schäfer, and Dominik Begerow. Ruhr-University<br />

Bochum, Geobotanik, ND03/174, 44780 Bochum, Germany. Alaska - Europe:<br />

Biogeography <strong>of</strong> arctic populations <strong>of</strong> Microbotryum<br />

Climatic changes over millions <strong>of</strong> years have shaped the vegetation <strong>of</strong> our<br />

earth. Especially plants <strong>of</strong> extreme habitats were heavily influenced and many<br />

species are extinct due to various ice ages in Europe. While there have been several<br />

studies on the population structure <strong>of</strong> higher plants <strong>of</strong> artic and alpine distribution,<br />

very little is known about the associated, host specific parasites. The aim<br />

<strong>of</strong> this study is the establishment <strong>of</strong> microsatellites for Microbotryum silenesacaulis<br />

based on low coverage genome sequencing and the analysis <strong>of</strong> populations<br />

from the Alps, Scandinavia and Alaska. Although the coverage is low,<br />

genome sequence data revealed hundreds <strong>of</strong> microsatellites <strong>of</strong> various sizes.<br />

While the number <strong>of</strong> bi- and trinucleotid repeats seems somehow limited, there<br />

are especially many longer motives <strong>of</strong> up to nine basepairs repeated many times.<br />

To differentiate neighbouring populations as well, we used highly variable satellites<br />

to analyse the structure <strong>of</strong> alpine and arctic populations. The link <strong>of</strong> M.<br />

silenes-acaulis collected one year ago during the MSA meeting in Fairbanks to<br />

European alpine or arctic populations will be discussed.<br />

Eyre, Catherine A, Melina Kozanitas, and Matteo Garbelotto. Forest Mycology<br />

and Pathology Laboratory, Department <strong>of</strong> Environmental Science, Policy and<br />

Management, University <strong>of</strong> California, Berkeley, CA 94720, USA. Population<br />

dynamics <strong>of</strong> aerial and terrestrial populations <strong>of</strong> Phytophthora ramorum in a<br />

California watershed under different climatic conditions<br />

In the attempt to better understand the epidemiology <strong>of</strong> Sudden Oak<br />

Death in California, we present the first combined genetic analysis <strong>of</strong> soil and leaf<br />

genotypes performed for a plant pathogen. Isolates <strong>of</strong> Phytophthora ramorum<br />

were collected from leaves and soil in six plots during a drought and an averagerainfall<br />

year, and analyzed using polymorphic SSRs. Leaf populations were resampled<br />

in different seasons during both years to provide information on how seasonal<br />

weather patterns may affect population dynamics. Confirmed SOD<br />

symptoms on leaves <strong>of</strong> the transmissive host Bay Laurel increased from 15 to<br />

39% between dry and wet conditions. Symptoms caused by other foliar pathogens<br />

were highest (69%) in dry conditions, suggesting that, while P. ramorum and<br />

other pathogens occupy the same niche, they are favored by different climatic<br />

conditions. Some foliar genotypes <strong>of</strong> P. ramorum were more abundant in wet than<br />

in dry conditions and persistent through time. Conversely, persistence <strong>of</strong> soil<br />

genotypes was limited. Soil and foliar populations were genetically distinct,<br />

(FST>0.15), but not segregated into different portions <strong>of</strong> a minimum spanning<br />

network, suggesting intermixing <strong>of</strong> the two. We surmise that the genetic structure<br />

between substrates results from genotype-specific differences in substrate adaptation<br />

rather than lack <strong>of</strong> gene flow; this hypothesis is supported by different genotype<br />

rank order abundances between soil and leaf populations. In climatic condi-<br />

tions unfavorable to the pathogen, genetic diversity increases both within and between<br />

sites, while in favorable conditions diversity decreases due to greater migration<br />

levels <strong>of</strong> some genotypes. Foliar genotypes can spread farther in wet<br />

years, and soil appears to be re-inoculated on a yearly basis. Cumulatively, our results<br />

indicate that leaves act as a relatively persistent source <strong>of</strong> inoculum, while<br />

soil is a sink that is genetically very diverse, but potentially inconsequential for the<br />

natural spread <strong>of</strong> the disease.<br />

Fisher, Karen E, Ricardo Reyes, David Lowry, and Robert W Roberson. School<br />

<strong>of</strong> Life Sciences, Arizona State University, Tempe, AZ 85287. Changes in hyphal<br />

growth and ultrastructure in two chitin synthase mutants <strong>of</strong> Neurospora<br />

crassa<br />

The primary fibrillar component <strong>of</strong> the fungal cell wall is chitin; a linear<br />

polymer <strong>of</strong> ß-1,4-N-acetylglucosamine. Chitin synthesis occurs at the cell surface,<br />

primarily at sites <strong>of</strong> growth, and is catalyzed by a trans-membrane enzyme known<br />

as chitin synthase (CHS). Seven CHSs have been identified in Neurospora crassa.<br />

Previous research utilizing GFP fusions <strong>of</strong> CHS-1, CHS-3 and CHS-6 has<br />

documented their accumulation in the core <strong>of</strong> the Spitzenkörper (Spk) in growing<br />

hyphae <strong>of</strong> N. crassa. The translocation <strong>of</strong> these enzymes, presumably by vesicle<br />

motility, towards the Spk from subapical locations has also been reported. Given<br />

the importance <strong>of</strong> chitin synthesis in fungal cell growth and morphogenesis, we<br />

have conducted a study <strong>of</strong> two chitin synthase deletion mutants in N. crassa:<br />

∆CHS-1 and ∆CHS-6. In this presentation, hyphal growth behavior will be described<br />

in these mutants relative to wild-type hyphae. In addition, ultrastructural<br />

analysis <strong>of</strong> cytoplasmic organization will be documented.<br />

Floudas, Dimitrios and David S Hibbett. Biology Department, Clark University,<br />

Worcester, MA, 01610. Addressing the polyphyletic origin <strong>of</strong> Phanerochaete<br />

(Polyporales, Basidiomycota) based on a multi-gene dataset: how many lineages<br />

exist?<br />

Phanerochaete Karst. is a saprotrophic basidiomycete genus <strong>of</strong> the Polyporales,<br />

which harbors white rot species that remove both lignin and carbohydrates<br />

from wood. One species in the genus, Phanerochaete chrysosporium, is a<br />

model system for the study <strong>of</strong> white rot biochemistry and its genome has been sequenced,<br />

revealing a wide range <strong>of</strong> genes related to lignocellulose degradation.<br />

Despite progress on deciphering the white rot mechanism <strong>of</strong> P. chrysosporium,<br />

we lack a deeper understanding <strong>of</strong> the phylogenetic relationships <strong>of</strong> the genus as<br />

a whole. The lack or simplicity <strong>of</strong> macro and micro-morphological characters <strong>of</strong><br />

the basidiocarps has rendered the identification <strong>of</strong> Phanerochaete species and the<br />

delimitation <strong>of</strong> the genus perplexing. Molecular studies, based mainly on ribosomal<br />

genes, have shown that the genus as traditionally delimited is <strong>of</strong> polyphyletic<br />

origin nested in Polyporales, but only a few Phanerochaete species have been<br />

incorporated in multi-locus phylogenetic studies. The fragmented phylogenetic<br />

picture for Phanerochaete hinders understanding <strong>of</strong> the morphological, biochemical<br />

and ecological evolution in the genus. We are sampling here multiple loci<br />

across Phanerochaete and related genera, as a part <strong>of</strong> the PolyPEET project, in an<br />

attempt to provide the first higher-level phylogeny <strong>of</strong> the genus, to assess its relationships<br />

with other genera in the Polyporales, and reassess the phylogenetic distribution<br />

<strong>of</strong> morphological characters. The preliminary results indicate that<br />

Phanerochaete s.l. species are found in at least 6 lineages in the phlebioid clade,<br />

highlighting the necessity for nomenclatural rearrangements in the genus.<br />

Gajdeczka, Michael T, Wenjun Li, Catharine B Kappauf, and Joseph Heitman.<br />

Department <strong>of</strong> Molecular Genetics and Microbiology, Duke University Medical<br />

Center, Durham, North Carolina, United States <strong>of</strong> <strong>America</strong>. Like a rock?: Lack<br />

<strong>of</strong> sequence variation in clinical isolates <strong>of</strong> the dermatophyte Trichophyton<br />

rubrum<br />

Trichophyton rubrum (Ascomycota) - the cause <strong>of</strong> athlete’s foot, ringworm<br />

and nail infections - is the most common dermatophyte affecting humans.<br />

Despite availability <strong>of</strong> an annotated genome, this microbe’s pathogenicity, epidemiology<br />

and genetics remain poorly understood due to an inability to conduct<br />

genetic crosses, lack <strong>of</strong> adequate sampling and apparent clonality. The only evidence<br />

<strong>of</strong> sexual mating in T. rubrum is indirect - a single genetic hybrid was<br />

formed between T. rubrum and Arthroderma simii, the teleomorph <strong>of</strong> a related<br />

species (T. mentagrophytes). T.rubrum lacks a known teleomorph, and although<br />

its genome contains mating type genes, only a single MAT allele (MAT1-1) is<br />

known. This study has three aims: (1) survey sequence variation in our 194 clinical<br />

samples using sequence-based and previously-developed variable number<br />

tandem repeat (VNTR) markers; (2) use paired-end Illumina sequencing to identify<br />

additional polymorphism within and among strains belonging to different<br />

VNTR types, and (3) search for evidence <strong>of</strong> (same-sex) mating using genetic and<br />

culture-based methods. First, we sequenced 7.5 kb <strong>of</strong> mainly non-coding regions<br />

located within 20 kb <strong>of</strong> the ABC1 and CAP59 genes in ten T. rubrum isolates collected<br />

over a span <strong>of</strong> ten years. These sequences revealed six singleton polymor-<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 17


phisms, which distinguished only half <strong>of</strong> the strains. We are now transitioning to<br />

VNTR genotyping and whole genome Illumina sequencing to characterize the diversity<br />

more completely. Lastly, in order to test for formation <strong>of</strong> ascomata (sexual<br />

structures), we will co-culture ten <strong>of</strong> our T. rubrum isolates and a T. tonsurans<br />

isolate (MAT1-2) in single, pair-wise and triple combinations on dermatophyte<br />

mating medium. Cultures with extensive hyphal coiling (indicative <strong>of</strong> early sexual<br />

development) will be dissected and examined for presence <strong>of</strong> ascomata or ascospores.<br />

Thus far, our results provide no evidence <strong>of</strong> sexual reproduction and reconfirm<br />

low sequence variation in T. rubrum clinical isolates.<br />

Garbelotto, Matteo M and Todd W Osmundson. Department <strong>of</strong> Environmental<br />

Science, Policy and Management, University <strong>of</strong> California, Berkeley, CA 94720,<br />

USA. Is there a “best” unit <strong>of</strong> conservation for valuable mushroom species?<br />

A scientific approach based on studies on matsutakes, morels and truffles<br />

In the presence <strong>of</strong> limited economic resources, an approach is needed to<br />

maximize the results <strong>of</strong> conservation efforts focused on valuable mushroom<br />

species. Although the genomic nature <strong>of</strong> desirable mushroom phenotypes is still<br />

mostly unknown, we propose that a biogeographical approach combined with<br />

population genetics studies provide a concrete way to enact mushroom conservation<br />

plans. In the absence <strong>of</strong> precise knowledge <strong>of</strong> why a specific population<br />

should be protected, we suggest that the unit <strong>of</strong> conservation should be that <strong>of</strong> the<br />

“metapopulation” within each broad geographic region. It was first suggested in<br />

the late 1990s that habitat fragmentation was a major driver <strong>of</strong> speciation in Matsutakes.<br />

That intuition was recently confirmed by formal analysis on the effects<br />

<strong>of</strong> forest fragmentation on fungal community structure. The logistical advantage<br />

<strong>of</strong> focusing on habitat conservation, rather than on individual fungal species, is<br />

obvious. We show here that for recently evolved species, distribution <strong>of</strong> species<br />

tracks the presence <strong>of</strong> continuous forest cover, and an “isolation-by-distance” approach<br />

can be used to identify valuable conservation units. However, we present<br />

four examples <strong>of</strong> exceptions that need to be taken into account: a) The presence<br />

<strong>of</strong> significant geographic barriers, such as high mountain ranges, may accelerate<br />

drift; b) Significant bottlenecks caused by geologically recent events like glaciations<br />

may have led to genetic isolation between populations surviving in different<br />

refugia; c) Natural ecological events such as fires may be triggers <strong>of</strong> speciation<br />

and <strong>of</strong> hybridization among mushrooms; and d) Forest management may accelerate<br />

genetic drift by favoring short rotations where gene flow among populations<br />

is hindered.<br />

Gause, Justin W and Merlin M White. Boise State University, Dept. <strong>of</strong> Biological<br />

Sciences, Boise, ID 83725. Entangled and nearly deceived in a web <strong>of</strong> gut<br />

fungi: two intermingled, dimorphic Smittium species from Idaho<br />

Since 2007 we have been prospecting for gut fungi near Boise, Idaho. The<br />

Boise River, which runs parallel to Boise State’s campus, sporadically has been a<br />

convenient and favorable system for aquatic insect collection, despite that it runs<br />

through this metropolitan area. Collections have included midges (Chironomidae),<br />

which are hosts <strong>of</strong> several genera <strong>of</strong> Harpellales, including Smittium. In<br />

2008, several collections <strong>of</strong> larval midges revealed an astonishing density <strong>of</strong><br />

hindgut dwelling gut fungi. Upon morphometric analysis <strong>of</strong> vouchered slides, two<br />

species <strong>of</strong> Smittium, both with dimorphic trichospores, were found to be coexisting<br />

inhabitants in the midge population. In addition to the dimorphic trichospores,<br />

zygospores were also recovered. Zygospores are <strong>of</strong>ten the missing “piece <strong>of</strong> the<br />

puzzle” for taxonomic considerations, and we were fortunate enough to have<br />

them within a few dissections. However, these specimens were remarkably well<br />

developed and presented intermingled thalli, demanding careful scrutiny to tease<br />

apart the key morphological aspects belonging to each. We present our data on the<br />

two species <strong>of</strong> Smittium and compare them with records <strong>of</strong> putatively closely related<br />

Smittium species, with suggestions for future efforts with these and other<br />

host populations in pursuit <strong>of</strong> gut fungi worldwide, perhaps in an area near you?<br />

Geiser, David M 1 , Alejandro P Rooney 2 , Todd J Ward 2 , Takayuki Aoki 3 ,<br />

Seogchan Kang 1 , Stephen A Rehner 4 , and Kerry O’Donnell 2 . 1 Department <strong>of</strong><br />

Plant Pathology, Penn State University, University Park, PA 16802, 2 USDA-<br />

ARS-NCAUR, 1815 N. University St., Peoria, IL 61604, 3 MAFF, NIAS, 2–1–2,<br />

Kannondai, Tsukuba, Ibaraki 305–8602, Japan, 4 Systematic Mycology and Microbiology<br />

Laboratory, USDA, ARS, Beltsville, MD 20705. Phylogenetically<br />

marking the limits <strong>of</strong> the genus Fusarium for post-Article 59 usage<br />

Fusarium (Hypocreales, Nectriaceae) is one <strong>of</strong> the most important and<br />

systematically challenging groups <strong>of</strong> mycotoxigenic, plant pathogenic and human<br />

pathogenic fungi. We conducted maximum likelihood (ML), maximum parsimony<br />

(MP) and Bayesian (B) analyses on partial nucleotide sequences <strong>of</strong> genes encoding<br />

the largest (RPB1) and second largest (RPB2) RNA polymerase II subunits<br />

on 94 fusaria to infer phylogenetic relationships within the genus and 20 <strong>of</strong><br />

its near relatives. Our analysis revealed that Cylindrocarpon spp. formed a basal<br />

monophyletic sister to a ‘terminal Fusarium clade’ (TFC sensu Gräfenhan et al.<br />

2011), which comprised 20 strongly informally named clades. Inclusion <strong>of</strong> the<br />

basal most divergences within the TFC received strong support only from the<br />

18 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Bayesian analysis (posterior probability (B-PP) = 0.99-1), with clades including<br />

F. ventricosum and F. dimerum forming the two earliest diverging lineages. An<br />

internode supporting the remaining TFC, however, was strongly supported by MP<br />

and ML boostrapping (ML-BS 100%, MP-BS 87%) and a B-PP <strong>of</strong> 1. This contrasts<br />

with the published analysis <strong>of</strong> a phylogeny based on RPB2 and acyl citrate<br />

lyase gene sequences, which did not yield significant support for this clade. Diversification<br />

time estimates were made to infer the origins <strong>of</strong> the genus and its<br />

clades. Because 1) the TFC includes all economically important fusaria, 2) virtually<br />

all <strong>of</strong> its members produce a recognizable Fusarium anamorph, 3) the Fusarium<br />

anamorph names are widely used compared to the associated teleomorphs;<br />

applied mycologists rarely if ever see the latter, and 4) it corresponds most closely<br />

to a fairly stable concept <strong>of</strong> the genus that has existed for nearly a century, herein<br />

we formally combine several teleomorph genera within the TFC and their types<br />

into Fusarium. The present study provides a robust framework to guide future<br />

comparative phylogenetic and phylogenomic analyses, including work to test and<br />

extend the hypotheses developed herein.<br />

Geml, Jozsef 1 , Eduardo R Nouhra 2 , Christian Y Wicaksono 1 , Nicolas Pastor 2 ,<br />

Lisandro Fernandez 2 , and Alejandra G Becerra 2 . 1 Nationaal Herbarium Nederland,<br />

Nederlands Centrum voor Biodiversiteit Naturalis, Universiteit Leiden, P.O.<br />

Box 9514, 2300RA Leiden, The Netherlands, 2 Instituto Multidisciplinario de Biología<br />

Vegetal (CONICET), Universidad Nacional de Córdoba, cc 495, cp 5000<br />

Córdoba, República Argentina. Mycota <strong>of</strong> the Andean Yungas forests: assessments<br />

<strong>of</strong> fungal biodiversity and habitat partitioning in a threatened ecosystem<br />

The Yungas, a system <strong>of</strong> subtropical montane forests on the eastern slopes<br />

<strong>of</strong> the Andes, reach their southern limit in northwestern Argentina. These forests<br />

are extremely diverse and, despite covering only 2% <strong>of</strong> the country’s area, they<br />

harbour about 50% <strong>of</strong> Argentina’s biodiversity. Unfortunately, the Yungas are<br />

among the ecosystems most threatened by anthropogenic pressure and climatic<br />

changes. Previous mycological works in the Yungas focused on wood-decay<br />

fungi (e.g., polypores) and mycorrhizae in Alnus acuminata cloud forests, while<br />

diverse Yungas communities still remain virtually unexplored. We carried out<br />

massively parallel pyrosequencing <strong>of</strong> ITS rDNA from soil samples to provide the<br />

first kingdom-wide fungal biodiversity assessment for the Yungas. Samples were<br />

taken in the three major forest types along an altitudinal gradient: the piedmont<br />

forest (400-700 m asl), the montane forest (700-1500 m asl), and the montane<br />

cloud forest (1500-3000 m asl). Using a 97% similarity cut-<strong>of</strong>f value, we delimited<br />

1839 fungal operational taxonomic units (OTUs) in total. The majority belonged<br />

to the phylum Ascomycota (52.49%), followed by Basidiomycota<br />

(18.85%), Glomeromycota (0.012%), and various zygomycete lineages (0.02%),<br />

while 24.91% showed highest similarity to other unidentified environmental sequences<br />

and could not be assigned to phylum. The distribution <strong>of</strong> the total 1839<br />

OTUs among the major altitudinal vegetation types were as follows: 909 OTUs<br />

in the piedmont forest, 826 in the montane forest and 929 in the montane cloud<br />

forest. Fungal communities were significantly different among all three forest<br />

types, with many OTUs showing strong habitat preference for a certain altitudinal<br />

zone. Our data <strong>of</strong>fers an unprecedented insight into the fungal biodiversity <strong>of</strong><br />

the Yungas and into the zonal changes in fungal community structure, with potential<br />

applications in conservation strategies to preserve the unique biodiversity<br />

<strong>of</strong> the Andean forests.<br />

Gonzalez, Maria C 1 , Anthony E Glenn 2 , Edmundo Rosique-Gil 3 , Kevin D<br />

Hyde 4 , and Richard T Hanlin 5 . 1 Departamento de Botánica, Instituto de Biología,<br />

Universidad Nacional Autónoma de México, Mexico City, DF 04510,<br />

Mexico, 2 Russell Research Center, Toxicology & Mycotoxin Research Unit,<br />

USDA, ARS, Athens, GA 30605, USA, 3 División Académica de Ciencias Biológicas,<br />

Universidad Autónoma Juárez de Tabasco, Villahermosa Tab 86039,<br />

Mexico, 4 5<br />

School <strong>of</strong> Science, Mae Fah Luang University, Chiang Rai, Thailand,<br />

Museum <strong>of</strong> Natural History Annex, University <strong>of</strong> Georgia, Bogart, GA 30622,<br />

USA. A new species <strong>of</strong> Aliquandostipite from an urban lagoon <strong>of</strong> Tabasco,<br />

Mexico<br />

During a continued survey <strong>of</strong> Ascomycota diversity from different lentic<br />

freshwater habitats, we recorded several interesting fungi from a lagoon located<br />

in the urbanized area <strong>of</strong> the Villahermosa City, capital <strong>of</strong> the State <strong>of</strong> Tabasco. An<br />

uncommon specimen was obtained using a submerged wood-baiting technique.<br />

This Mexican microorganism shows some <strong>of</strong> the typical characteristics <strong>of</strong> Aliquandostipite<br />

type species, A. khaoyaiensis i.e. subglobose, ostiolate ascomata; 8spored,<br />

bitunicate, asci; bi-celled, multiguttulate, sheated ascospores and pigmented<br />

hyphae up to 40 µm wide. However, in addition, it has other relevant<br />

features such as the presence <strong>of</strong> very long, subcylindrical, contorted, branched, ascomatal<br />

neck that when young is covered with hyaline, straight, attenuated, 20-50<br />

µm x 2-4 µm hairs, a characteristic is not described in Aliquandostipite species.<br />

Other characteristics that the new fungus exhibits include different average size <strong>of</strong><br />

mature ascomata, asci and ascospores, the distinctly bitunicate, thin-walled, deli-<br />

Continued on following page


quescent asci and ascospores deeply constricted at the septum that separate into<br />

two part spores. The analysis <strong>of</strong> morphological set <strong>of</strong> characters and ITS rDNA<br />

sequence data confirmed this Mexican lignicolous fungus belongs to Aliquandostipite,<br />

but it is genetically different from all the other described species. Therefore,<br />

this fungus is proposed as a new species and is described and illustrated. The<br />

high frequency <strong>of</strong> occurrence <strong>of</strong> this ascomycete at heavily polluted tropical urban<br />

lagoon possibly indicates a successful adaptation to that habitat.<br />

Goodwin, Stephen B. USDA-ARS, Department <strong>of</strong> Botany and Plant Pathology,<br />

915 West State Street, Purdue University, West Lafayette, IN 47907-2054. Evolution<br />

<strong>of</strong> plant pathogenesis within Dothideomycetes<br />

The Dothideomycetes is the largest class <strong>of</strong> fungi, occurring on every continent<br />

and in very diverse habitats. In addition to many plant pathogens in the orders<br />

Capnodiales and Pleosporales, which collectively cause economically important<br />

diseases on almost every major crop, Dothideomycetes can endure<br />

extremes <strong>of</strong> heat or cold, high solar radiation, desiccation, live in fresh- or saltwater<br />

habitats, persist as lichens or can form mycorrhizal associations with plants.<br />

To better understand plant pathogenicity and the genetic basis for this huge ecological<br />

diversity, the genomes <strong>of</strong> more than 20 Dothideomycetes have been sequenced<br />

or are in progress, ranging in size from 22 Mb for the extremophile Baudoinia<br />

compniacensis to 74 Mb for the banana pathogen Mycosphaerella fijiensis.<br />

Genome expansion seems to have occurred primarily by proliferation <strong>of</strong> retrotransposons,<br />

but other mechanisms were noted including the first evidence for<br />

transduplication by DNA transposons in fungi. An unusual component <strong>of</strong> the M.<br />

graminicola genome was a “dispensome” <strong>of</strong> eight chromosomes that can be lost<br />

with no obvious effect on fitness, appears to have originated by horizontal transfer<br />

more than 10,000 years ago, and is highly variable in field populations. Some<br />

other Dothideomycetes genomes have possible dispensable chromosomes that are<br />

unrelated to those in M. graminicola. Functional characterization <strong>of</strong> the genomes<br />

revealed that many species <strong>of</strong> Mycosphaerella have reduced sets <strong>of</strong> genes for cell<br />

wall-degrading enzymes and other proteins, possibly as an adaptation to avoid detection<br />

by their hosts to facilitate stealth pathogenicity. Effectors identified in the<br />

genome <strong>of</strong> M. fijiensis by similarity to those in the tomato pathogen Cladosporium<br />

fulvum could be recognized by the heterologous resistance gene to elicit a defense<br />

response. The extensive genomic resources for the Dothideomycetes provide<br />

tools for a comprehensive understanding <strong>of</strong> the genetics and evolution <strong>of</strong><br />

adaptations to very diverse environments and modes <strong>of</strong> nutrition.<br />

Gossage, Zachary T 1 , Sagar Yeraballi 1 , Katherine Suding 2 , Robert Sinsabaugh<br />

3 , and Andrea Porras-Alfaro 1 . 1 Biological Sciences, Western Illinois University,<br />

Macomb, IL, 2 Department <strong>of</strong> Environmental Science, Policy & Management,<br />

University <strong>of</strong> California Berkeley, CA, 3 Department <strong>of</strong> Biology, University<br />

<strong>of</strong> New Mexico, Albuquerque, NM. Root-associated fungi <strong>of</strong> Geum rossii and<br />

Deschampsia cespitosa at the Niwot LTER site<br />

Plant-associated fungi are known to play an important role in plant productivity<br />

and ecosystem sustainability. Endophytic fungal diversity in the alpine<br />

tundra is unknown for the majority <strong>of</strong> plant species. The main goal <strong>of</strong> this study<br />

was to identify and describe fungal symbionts associated with two alpine tundra<br />

co-dominant plants: Geum rossii and Deschampsia cespitosa. Plants were collected<br />

at Niwot Ridge long-term nitrogen fertilization experiments in Colorado.<br />

Forty-two plants were collected from four different treatments: control, nitrogen<br />

fertilized, D. cespitosa removal with nitrogen fertilization, and D. cespitosa removal<br />

with no nitrogen addition in 2008 and additional plants were collected from<br />

control plots in 2010. Roots were harvested from each treatment and stained to determine<br />

fungal colonization. Roots were surface sterilized and plated on malt extract<br />

agar with antibiotics. One hundred and ten pure cultures <strong>of</strong> endophytic fungi<br />

were isolated and identified by sequencing the internal transcribed spacer rDNA.<br />

Analysis <strong>of</strong> the sequences at 97% similarity yielded 25 contigs and 24 unique sequences.<br />

The majority <strong>of</strong> the isolates, 91%, belong to the phylum Ascomycota<br />

and the remaining to Basidiomycota and other undetermined phyla. Dominant orders<br />

included Helotiales (23%), Hypocreales (23%) and Eurotiales (21%). Within<br />

Helotiales, dominant taxa were closely related to Phialocephala fortinii and<br />

Cryptosporiopsis ericae. Within Hypocreales some common taxa included<br />

Tolypocladium and Beauveria. Dominant taxa within Eurotiales included isolates<br />

related to Penicillium sp. and Aspergillus sp. Endophytic communities showed<br />

very little plant specificity suggesting that most taxa are adapted to colonize both<br />

<strong>of</strong> these dominant species.<br />

Grigoriev, Igor. US DOE Joint Genome Institute, 2800 Mitchell Dr, Walnut<br />

Creek, CA 94598. Fueling the future with fungal genomics<br />

Until recently, genomic diversity <strong>of</strong> Fungi was biased towards ascomycetes<br />

<strong>of</strong> medical importance. To correct this bias the Fungal Genomics Program<br />

(jgi.doe.gov/fungi) <strong>of</strong> the US Department <strong>of</strong> Energy (DOE) Joint Genome<br />

Institute (JGI) has partnered with international scientific community to explore<br />

fungal diversity in several large scale genomics initiatives aligned with the 2010<br />

Grand Challenges for Biological and Environmental Research: a long term vision.<br />

The first initiative, the Genomic Encyclopedia <strong>of</strong> Fungi, is focused on diversity<br />

among DOE relevant fungi in the areas <strong>of</strong> plant health, to explore the interactions<br />

<strong>of</strong> bioenergy crop species with symbionts and pathogens, and <strong>of</strong> biorefinery, to<br />

catalog industrially relevant genes, pathways, and hosts for bioenergy applications.<br />

The second initiative, the 1000 Fungal Genomes project, is aimed to explore<br />

fungal diversity across the Fungal Tree <strong>of</strong> Life in order to provide references<br />

for research on plant-microbe interactions and environmental metagenomics. The<br />

third initiative is to support functional studies <strong>of</strong> fungal systems <strong>of</strong> varying complexity:<br />

from new model organisms to metagenomes <strong>of</strong> complex communities.<br />

Open to all scientists around the world, these initiatives result in massive amounts<br />

<strong>of</strong> genomic information integrated with analytical tools and community-driven<br />

experiments. Several examples produced by these initiatives will illustrate how<br />

genomic data and different sampling strategies can improve our understanding <strong>of</strong><br />

fungal diversity, interactions, and evolution.<br />

Grube, Martin 1 , Cene Gostincar 2 , and Lucia Muggia 1 . 1 Institute <strong>of</strong> Plant Sciences,<br />

Karl-Franzens-University Graz, Graz, Austria, 2 Department <strong>of</strong> Biology,<br />

Biotechnical Faculty, University <strong>of</strong> Ljubljana, Ljubljana, Slovenia. Evolutionary<br />

aspects <strong>of</strong> oligotrophism and extremotolerance in Ascomycota<br />

As widely distributed colonizers <strong>of</strong> rocks, black meristematic fungi endure<br />

temperature extremes, desiccation and usually a low supply <strong>of</strong> nutrients. Certain<br />

traits, such as phenotypic plasticity and melanized cell walls have been suggested<br />

as important factors <strong>of</strong> their polyextremotolerance. Some <strong>of</strong> their adaptive<br />

abilities, e.g. growth at elevated temperatures, may also be considered as health<br />

concerns, especially since black fungi can <strong>of</strong>ten establish themselves in manmade<br />

habitats. Facultative lichen-like associations <strong>of</strong> black fungi with algae could<br />

contribute to the persistence in poor natural habitats on rocks. We detected several<br />

cases <strong>of</strong> such associations. Phylogenetic analyses already suggested an ancestral<br />

proximity <strong>of</strong> oligotrophic black fungi and certain lineages <strong>of</strong> lichen-forming<br />

fungi. We included further lichenized and lichenicolous species in phylogenetic<br />

analyses that seem to link certain groups <strong>of</strong> lichens with black fungi. We believe<br />

that the upcoming information on genomes from black fungi and comparative genomics<br />

will provide us with new insights into the pre-adaptations for extremotolerance<br />

and oligotrophism <strong>of</strong> black fungal lineages. This knowledge may also be<br />

used to predict the future evolutionary potentials <strong>of</strong> certain fungal lineages.<br />

Gryganskyi, Andrii P 1,5 , Richard A Humber 2 , Matthew E Smith 3 , Jolanta Miadlikovska<br />

1 , Steven Wu 1 , Kerstin Voigt 4 , Iryna M Anishchenko 5 , Oleksandr V<br />

Savytskyi 6 , and Rytas Vilgalys 1 . 1 Duke University, Department <strong>of</strong> Biology,<br />

Durham, NC 27708-90338, USA, 2 USDA-ARS BioIPM Research, RW Holley<br />

Center for Agriculture & Health 538 Tower Road, Ithaca, NY 14853, USA, 3 Department<br />

4<br />

<strong>of</strong> Plant Pathology, University <strong>of</strong> Florida, Gainesville, FL 32611, USA,<br />

Friedrich-Schiller University <strong>of</strong> Jena, Institute <strong>of</strong> Microbiology, 25 Neugasse,<br />

Jena 07743, Germany, 5 M. Kholodny Institute <strong>of</strong> Botany, National Academy <strong>of</strong><br />

Sciences <strong>of</strong> Ukraine, 2 Tereshchenkivska St., Kyiv 01601, Ukraine, 6 Institute <strong>of</strong><br />

Molecular Biology and Genetics, National Academy <strong>of</strong> Sciences <strong>of</strong> Ukraine, 150<br />

Acad. Zabolotnogo St., Kyiv 03680, Ukraine. Molecular phylogeny <strong>of</strong> the Entomophthoromycota.<br />

The Entomophthoromycota, a recently recognized new phylum, is a ubiquitous<br />

group <strong>of</strong> fungal pathogens <strong>of</strong> a wide variety <strong>of</strong> economically important insect<br />

pests, and other soil invertebrates. This group also includes a small number<br />

<strong>of</strong> parasites <strong>of</strong> reptiles, vertebrates (including humans), macromycetes, fern gametophytes,<br />

and desmid algae, as well as some saprobes. Here we report recent<br />

studies to resolve the phylogenetic relationships within the Entomophthoromycota<br />

and to place this group reliably among other basal fungal lineages. Bayesian Interference<br />

and Maximum Likelihood analyses <strong>of</strong> four loci (nuclear 18S and 28S<br />

rDNA, mitochondrial 16S, and the protein-coding RPB2) as well as non-molecular<br />

data consistently and unambiguously identify 31 taxa <strong>of</strong> Entomophthoromycota<br />

as a monophyletic group distinct from other Zygomycota and flagellate fungi.<br />

Using the constraints <strong>of</strong> our multi-gene dataset we constructed the most comprehensive<br />

rDNA phylogeny yet available for Entomophthoromycota. The taxa studied<br />

here belong to five distinct, well-supported lineages. The Basidiobolus clade<br />

is the earliest diverging lineage. The Conidiobolus lineage is a paraphyletic grade<br />

<strong>of</strong> trophically diverse species that include saprobes, insect pathogens, and facultative<br />

human pathogens. Three well supported lineages in the exclusively entomopathogenic<br />

Entomophthoraceae center around the genera Batkoa, Entomophthora<br />

and Zoophthora, although several genera within this latter crown clade are<br />

resolved as non-monophyletic. Ancestral state reconstruction suggests that the ancestor<br />

<strong>of</strong> all Entomophthoromycota was morphologically similar to species <strong>of</strong><br />

Conidiobolus. Analyses using strict, relaxed, and local molecular clock models<br />

documented highly variable DNA substitution rates among lineages <strong>of</strong> Entomophthoromycota.<br />

Despite the complications caused by different rates <strong>of</strong> molecular<br />

evolution among lineages, our dating analysis indicates that the Entomophthoromycota<br />

originated 405±90 million years ago. We suggest that<br />

entomopathogenic lineages in Entomophthoraceae probably evolved from sapro-<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 19


ic or facultatively pathogenic ancestors during (or shortly after) the evolutionary<br />

radiation <strong>of</strong> the arthropods.<br />

Hameed, Khalid 1 , Gregory Bonito 1 , Michael Gajdeczka 1 , Christopher Schadt 2 ,<br />

and Rytas Vilgalys 1 . 1 Biology Department, Duke University, Durham NC<br />

27708, 2 Oak Ridge National Laboratory, Department <strong>of</strong> Energy, Oak Ridge TN<br />

37831. Fungal communities <strong>of</strong> Populus - comparisons between culture-based<br />

and culture independent approaches<br />

Populus is an economically important genus <strong>of</strong> woodland and riparian<br />

trees that forms both arbuscular and ectomycorrhizas. Populus also harbors a<br />

wide-diversity <strong>of</strong> root endophytes but this diversity is still largely uncharacterized.<br />

To address this field and greenhouse studies were conducted on different Populus<br />

species and genotypes. Fungal community diversity in roots was determined with<br />

culture-independent 454 pyrosequencing <strong>of</strong> phylogenetically informative fragments<br />

<strong>of</strong> genomic rDNA (ITS, LSU, SSU). Companion culture-based studies <strong>of</strong><br />

Populus root associates were carried out to: 1) describe the culturable diversity <strong>of</strong><br />

Populus root associates; 2) determine discrepancies between culture-based and<br />

culture-independent methods; 3) obtain a culture library <strong>of</strong> fungal root associates<br />

<strong>of</strong> Populus for future experiments and genome sequencing. Using sterile technique<br />

and selective isolation medias, we isolated >300 cultures from individual<br />

surface sterilized root tips <strong>of</strong> Populus deltoides collected from our field site and<br />

greenhouse experiments. We also isolated fungi from roots <strong>of</strong> Pinus taeda (loblolly<br />

pine) and Quercus alba (white oak) for comparison. Pure cultures were sequenced<br />

at the ITS and LSU. DNA sequences were identified by comparing them<br />

to accessions in the NCBI nucleotide database and to sequences assembled from<br />

pyrosequencing data. We estimate our cultures represent ~65 OTUs based on<br />

97% similarity <strong>of</strong> the ITS. Despite using benomyl in our media to select for basidiomycetes,<br />

the majority <strong>of</strong> OTUs from sequenced isolates were ascomycetes.<br />

We obtained cultures from a number <strong>of</strong> the fungi that were among the most abundant<br />

sequence type in our 454 datasets (e.g., Atractiellales spp., Ilionectria spp.,<br />

Mortieriella spp, and Thelephoraceae sp.). Further, fruitbodies <strong>of</strong> Laccaria were<br />

found under a number <strong>of</strong> trap-plants and this species has been obtained in pure<br />

culture. These isolates will be valuable for further studies on Populus-microbial<br />

relations.<br />

Hansen, Karen and Ibai Olariaga. Swedish Museum <strong>of</strong> Natural History, Department<br />

<strong>of</strong> Cryptogamic Botany, P.O. Box 50007, SE-104 05 Stockholm, Sweden.<br />

A multigene phylogenetic assessment <strong>of</strong> species boundaries in Otidea<br />

Species <strong>of</strong> Otidea are among the more conspicuous Pyronemataceae<br />

(Pezizomycetes), producing large, typically ear-shaped apothecia. Otidea is restricted<br />

to the Northern Hemisphere, with the highest diversity in boreal-temperate<br />

forests, and is ectomycorrhizal. The genus is monophyletic and easy to recognize,<br />

but species delimitation is highly controversial. To resolve species<br />

boundaries and relationships within Otidea, we provide a multigene (LSU rDNA,<br />

RPB1, RPB2 and EF1; 4908 bp), worldwide phylogeny <strong>of</strong> Otidea and comparative<br />

morphological data. Two main lineages are supported within Otidea: I) O.<br />

alutacea, O. apophysata, O. daliensis, O. platyspora, and O. subterranea, and II)<br />

the rest <strong>of</strong> Otidea, which encompasses several sub-lineages. Otidea papillata and<br />

a clade <strong>of</strong> O. leporina and Otidea sp., are identified as successive sister lineages<br />

to the rest <strong>of</strong> lineage II. No unique morphological characters distinguish the two<br />

main lineages. The sub-lineages are supported by a combination <strong>of</strong> morphological<br />

characters, such as the apothecia shape and color, spore shape and ornamentation,<br />

paraphyses shape and content, and outer excipulum features. Our results,<br />

including sequences <strong>of</strong> type collections, show clear species limits, except within<br />

two species complexes: the O. alutacea-cochleata and O. phlebophora-rainierensis<br />

complexes. A number <strong>of</strong> Otidea names currently in use are shown to be synonyms.<br />

Several names have been interpreted differently in Europe, North <strong>America</strong><br />

and Asia. At the same time some species seem to show continental endemism.<br />

Five new species have been discovered. The ITS region is too variable to align<br />

across Otidea, but appears to be useful for delimiting species; the ITS region is<br />

highly variable between species, but conserved within. To delimit species within<br />

the two species complexes we will perform separate analyses <strong>of</strong> these, including<br />

also the ITS region, and explore various analytical approaches for independent<br />

corroboration <strong>of</strong> the species identified. We estimate that Otidea comprises at least<br />

35 species.<br />

Hart, Andrew M and Thomas J Volk. Department <strong>of</strong> Biology, University <strong>of</strong><br />

Wisconsin, La Crosse, WI 54601, USA. Trametes versicolor: A tremendously<br />

versatile fungus<br />

Since microbes are generally seen in a negative public light, it is refreshing<br />

to take a look at one that is not only beneficial, but has the potential to serve<br />

multiple human interests. Trametes versicolor is a white rot polypore commonly<br />

referred to as the “turkey tail” mushroom. It is named for and characterized by the<br />

many concentric zones <strong>of</strong> color on the fruiting body. It is important to note that<br />

due to its prevalence in the northern forests, the turkey tail is an important saprophyte,<br />

but this might be the least useful quality <strong>of</strong> T. versicolor. The same versa-<br />

20 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

tile enzyme (laccase) that degrades lignin is key to several <strong>of</strong> T. versicolor’s applications.<br />

Laccases are used in biopulping and the bioremediation <strong>of</strong> a variety <strong>of</strong><br />

environmental contaminants. T. versicolor produces this valuable enzyme, and<br />

also gives us polysaccharide-K (PSK). Polysaccharide-K is anecdotally proven to<br />

function as an adjuvant (immunostimulatory). Polysaccharide-K is thought to enhance<br />

immune response to cancer cells. In addition to medical and environmental<br />

uses, T. versicolor is also used for its pigments. The pigments increase the<br />

wood’s monetary value and in turn are prized by woodworkers and artists. The inoculated<br />

wood forms streaks <strong>of</strong> color, called spalting. T. versicolor pigments are<br />

also used to dye fabrics such as wool. The multi-colored velvety fruiting body is<br />

used to make art <strong>of</strong> all kinds. These are the uses we’ve found for the turkey tail so<br />

far. As more research is done we will see the common fungus even more. This<br />

poster is applicable for education in courses related to Mycology, Medicine,<br />

Bioremediation and Biodegradation, Environmental Toxicology, Ecology, Biochemistry,<br />

and Biology.<br />

Hayward, Jeremy and Thomas R Horton. Department <strong>of</strong> Environmental and<br />

Forest Biology, Syracuse University, Syracuse, NY 13210. Thelephoroid symbionts<br />

<strong>of</strong> Pisoniae are limited to a small group <strong>of</strong> closely related taxa<br />

throughout the range <strong>of</strong> the plant tribe<br />

While several studies have demonstrated greater ectomycorrhizal (ECM)<br />

specificity in the tropics than in the temperate zone, little is known about the patterns<br />

underlying this specificity. Here, we present preliminary results from a study<br />

investigating the structure <strong>of</strong> specificity between members <strong>of</strong> the Nyctaginaceae<br />

and the ECM /tomentella-thelephora lineage. Within the Nyctaginaceae ectomycorrhizal<br />

plants are restricted to the genera Neea, Pisonia and Guapira, all members<br />

<strong>of</strong> the tribe Pisoniae. Following a previous observation that two Pisoniae<br />

species in opposite hemispheres associated with closely related thelephoroids, we<br />

hypothesized that members <strong>of</strong> the Pisoniae, whatever their geographic context,<br />

would also associate with this clade. We sampled ectomycorrhizas from members<br />

<strong>of</strong> the Pisoniae growing in the Pacific and on the Caribbean islands <strong>of</strong> Dominica,<br />

Vieques and Puerto Rico. We used molecular barcoding to identify ECM fungi<br />

and analyzed their phylogenetic placement using Bayesian and Maximum-Likelihood<br />

techniques. We also undertook a literature review to identify fungi previously<br />

detected associating with members <strong>of</strong> the Pisoniae. We found that at the familial<br />

level, the ectomycorrhizal specificity <strong>of</strong> the Pisoniae is low: species<br />

associate with members <strong>of</strong> at least five ECM fungal families. In contrast to this<br />

pattern <strong>of</strong> family-level generalism, those thelephoroid ECM fungi associating<br />

with the Pisoniae cluster strongly within a single lineage, implying narrow specificity<br />

between the Pisoniae and their thelephoroid symbionts at a subgeneric level.<br />

This clustering pattern is well supported in nonparametric tests <strong>of</strong> phylogenetic<br />

signal. This pattern <strong>of</strong> specificity holds true in the Pacific, in the Caribbean and in<br />

South <strong>America</strong>; it also holds true for representatives <strong>of</strong> all ectomycorrhizal plant<br />

genera within the Pisoniae. Based on these results, we hypothesize that phylogenetic<br />

placement may matter more than location in determining associations<br />

among tropical trees displaying ECM specificity. This hypothesis should be tested<br />

in associations between other tropical ECM plants and fungi.<br />

Healy, Rosanne A 1 , Gregory M Bonito 2 , Matthew E Smith 3 , Gonzalo G Guevara<br />

4 , Jonathon Frank 5 , Darlene Southworth 5 , James M Trappe 6 , and David J<br />

Mclaughlin 1 . 1 Department <strong>of</strong> Plant Biology, University <strong>of</strong> Minnesota, St. Paul,<br />

MN 55108, USA, 2 Department <strong>of</strong> Biology, Duke University, Durham, NC<br />

27708, USA, 3 Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Florida, Gainesville<br />

FL, USA, 4 Instituto Tecnológico de Ciudad Victoria, Biology, Ciudad Victoria,<br />

Mexico, 5 Department <strong>of</strong> Biology, Southern Oregon University, Ashland, OR<br />

97520, 6 Department <strong>of</strong> Forest Ecosystems and <strong>Society</strong>, Oregon State University,<br />

Corvalis, OR, USA. What is a cup fungus doing fruiting from a truffle tree?<br />

The case <strong>of</strong> the truffle-cup fungus lineage /pachyphloeus-scabropezia<br />

The ectomycorrhizal genera Pachyphloeus, a truffle, and Scabropezia, a<br />

cup fungus, are found primarily in broad-leaved forests <strong>of</strong> the Northern Hemisphere.<br />

Respectively they have twelve (plus two varieties) and three accepted described<br />

species. Molecular phylogenetic analyses have placed Scabropezia within<br />

Pachyphloeus. Previous inferences <strong>of</strong> truffle evolution in the Pezizales suggest<br />

that fruitbody evolution proceeds from cup fungus form to truffle form with no<br />

known reversals. Current data is unclear on whether Scabropezia branches from<br />

the base <strong>of</strong> the Pachyphloeus lineage, or is nested within Pachyphloeus, making<br />

it paraphyletic. The latter scenario may imply a reversal <strong>of</strong> form from truffle to<br />

cup fungus. Taxonomic confusion in Pachyphloeus from poorly understood type<br />

descriptions has resulted in expansion beyond original species limits, which in<br />

turn has led to confusion regarding the relationship <strong>of</strong> Scabropezia to Pachyphloeus.<br />

The goals <strong>of</strong> this study were to revise the /pachyphloeus-scabropezia lineage<br />

using multigene analyses from sporocarp and anamorph DNA sequences,<br />

and to infer the relationship <strong>of</strong> Scabropezia to Pachyphloeus. The ITS rDNA was<br />

sequenced from representative specimens from across the Northern Hemisphere,<br />

and then sorted into phylotypes based on 96% similarity. The LSU, RPB1 and<br />

Continued on following page


RPB2 loci were sequenced from selected specimens to better resolve relationships<br />

within the lineage. Morphological data was mapped onto one <strong>of</strong> the best-supported<br />

maximum likelihood phylograms to identify phylogenetically informative features.<br />

Preliminary results suggest that 1) the Scabropezia clade branches from the<br />

base <strong>of</strong> the /pachyphloeus lineage, but includes P. austro-oregonensis, 2) there are<br />

at least 25 species <strong>of</strong> Pachyphloeus and four species <strong>of</strong> Scabropezia, 3) P. lateritius<br />

and P. macrosporus are not part <strong>of</strong> the lineage, and 4) anamorph color, sporocarp<br />

color, ascus shape, and spore wall ornamentation are congruent with molecular<br />

data.<br />

Heidmarsson, Starri 1 , Sergio Perez-Ortega 2 , Holger Thüs 3 , Cecile Gueidan 3 ,<br />

Asuncíon de los Ríos 2 , and Francois Lutzoni 4 . 1 Icelandic Institute <strong>of</strong> Natural History,<br />

Borgir Nordurslod, IS-600 Akureyri, Iceland, 2 Departamento de Biología<br />

Ambiental, Museo Nacional de Ciencias Naturales, Madrid, Spain, 3 Botany, Natural<br />

History Museum, London, United Kingdom, 4 Biology, Duke University,<br />

Durham, United States. Diversity and phylogeny <strong>of</strong> marine and freshwater<br />

Verrucariaceae<br />

Verrucariaceae consists mainly <strong>of</strong> lichenized members, although the<br />

lichenicolous lifestyle can also be found among members <strong>of</strong> this family. The family<br />

consists <strong>of</strong> crustose, squamulose, foliose and even subfruticose taxa. Within<br />

Verrucaria s. lat., delimitations <strong>of</strong> taxa has been problematic because only few<br />

phenotypic characters have been useful to infer relationships. Recent DNA-based<br />

phylogenetic analyses across this family have already resulted in the description<br />

<strong>of</strong> a genus comprising only marine members <strong>of</strong> Verrucaria s. lat. (Wahlenbergiella)<br />

and another genus which contains taxa that either grow by freshwater<br />

or in marine habitat (Hydropunctaria). Recently it was shown that Mastodia,<br />

which forms symbiosis with a leafy green alga, is part <strong>of</strong> the Wahlenbergiella<br />

clade. We will present the diversity <strong>of</strong> aquatic Verrucaria s. Lat. in Iceland, from<br />

both marine and freshwater habitats. Phylogenetic analyses <strong>of</strong> the genera Hydropunctaria<br />

and Wahlenbergiella were conducted, mainly based on Icelandic<br />

material, within a broader phylogenetic framework. These analyses revealed that<br />

the strictly marine clade Wahlenbergiella consists <strong>of</strong> at least two, fairly distinct<br />

clades. One <strong>of</strong> these clades comprises members <strong>of</strong> Wahlenbergiella, such as W.<br />

mucosa, W. striatula, W. Tavaresiae, and Verrucaria ceuthocarpa. The other<br />

clade consists <strong>of</strong> Mastodia tesselata and several taxa, which have been only collected<br />

in the southern hemisphere, so far.<br />

Heimdal, Rune 1 , Marie L Davey 1,2 , and Håvard Kauserud 1 . 1 Molecular Evolution<br />

Research Group, Department <strong>of</strong> Biology, University <strong>of</strong> Oslo, N-0316 Oslo,<br />

Norway, 2 Department <strong>of</strong> Ecology and Natural Resource Management, Norwegian<br />

University <strong>of</strong> Life Sciences, N-1432 Ås, Norway. Host and tissue specificities<br />

in Galerina and Mycena species associated with boreal forest mosses<br />

revealed by data-mining <strong>of</strong> amplicon pyrosequencing fungal inventories<br />

The agaricoid basidiomycete genera Galerina and Mycena are both common<br />

saprophytes in the boreal forest that are frequently reported in ‘mossy habitats’<br />

or in the presence <strong>of</strong> mosses. ITS2 Sequences with BLAST matches to Galerina<br />

and Mycena were extracted from three large scale, amplicon<br />

pyrosequencing-based inventories <strong>of</strong> moss-associated fungi in southern Norway,<br />

and used for investigations into the genetic diversity <strong>of</strong> both groups and their ecological<br />

preferences. In both genera, genetic diversity was high relative to reported<br />

alpha diversity in those habitats. A relatively small number <strong>of</strong> OTUs occurred<br />

with high frequency and abundance, and in addition, a larger number <strong>of</strong> taxa were<br />

rare and occurred at very low frequencies and abundances. Among those common<br />

taxa, most OTUs showed a distinct preference for senescent moss tissues, supporting<br />

the idea that these fungi are decomposers <strong>of</strong> bryophytes. However, a number<br />

<strong>of</strong> OTUs also occurred with high frequency and abundance in photosynthetic<br />

tissues, suggesting some species have a facility for colonization and persistence in<br />

living moss tissues. In addition, some OTU’s showed a distinct predilection for<br />

particular moss hosts, suggesting that identifying the moss associated with fruitbodies<br />

<strong>of</strong> these fungi may prove useful in the identification <strong>of</strong> the fungus.<br />

Henkel, Terry W 1 , Michael A Castellano 2 , Hannah T Reynolds 3 , M Catherine<br />

Aime 4 , Steven L Miller 5 , and Matthew E Smith 6 . 1 Department <strong>of</strong> Biological Sciences,<br />

Humboldt State University, Arcata, CA 95521, 2 USDA, Forest Service,<br />

Northern Research Station, 3200 Jefferson Way, Corvallis, OR 97331, 3 Department<br />

<strong>of</strong> Biology, University <strong>of</strong> Akron, Akron OH 44023, 4 Department <strong>of</strong> Plant<br />

Pathology and Crop Physiology, Louisiana State University Agricultural Center,<br />

Baton Rouge, LA 70803, USA, 5 Department <strong>of</strong> Botany, University <strong>of</strong> Wyoming,<br />

Laramie, WY 82071, 6 Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Florida,<br />

Gainesville FL 32611. Ectomycorrhizal ascomycetes from the Guiana Shield:<br />

Elaphomyces and Pseudotulostoma<br />

The hypogeous false truffle genus Elaphomyces Fries (Elaphomycetaceae,<br />

Eurotiales, Ascomycota) contains ectomycorrhizal (ECM) fungi with a<br />

gleba <strong>of</strong> powdery, dark ascospores, a thick peridium, and large ascomata relative<br />

to other members <strong>of</strong> the Eurotiales. Of the ~67 species <strong>of</strong> Elaphomyces currently<br />

described worldwide, many are known from north temperate forests with ECM<br />

host plants in the Pinaceae, Fagaceae, and Betulaceae. Nearly a century ago<br />

Cooke and Rodway provided the first reports <strong>of</strong> Elaphomyces from the Southern<br />

Hemisphere (Australia), but erroneously assigned European names to their specimens.<br />

Recent efforts by Castellano et al. have uncovered at least 16 new<br />

Elaphomyces species from Australia and New Zealand. While Corner and Hawker<br />

in 1955 were the first to publish new Elaphomyces species from the lowland<br />

tropics in association with Dipterocarpaceae in Singapore, few other tropical<br />

species have been described. The recently discovered Elaphomyces digitatus ined.<br />

and Elaphomyces compleximurus ined. from Guyana constitute the first records<br />

for the genus for the lowland South <strong>America</strong>n tropics. These fungi occur in rainforests<br />

dominated by ECM trees <strong>of</strong> the genus Dicymbe (Fabaceae subfam. Caesalpinioideae)<br />

and constitute the first records <strong>of</strong> Elaphomyces in association with<br />

leguminous host plants. Pseudotulostoma volvata, a stipitate, volvate member <strong>of</strong><br />

the Elaphomycetaceae, is common in the same forests. So far, these fungi are the<br />

only ECM ascomycetes known from this lowland tropical ecosystem, despite the<br />

fact that more than 200 species <strong>of</strong> ECM basidiomycetes are present. We will discuss<br />

the morphology and ecology <strong>of</strong> these taxa, the evolution <strong>of</strong> ascoma form in<br />

the Elaphomycetaceae, and “missing” Elaphomyces diversity uncovered by rootbased<br />

sequencing studies.<br />

Herr, Joshua R. 321 Forest Resources Building, School <strong>of</strong> Ecosystem Sciences,<br />

The Pennsylvania State University, University Park, PA 16801. Fungal diversity<br />

associated with differing monoculture forest types measured by next-generation<br />

metabarcoding.<br />

It has long been understood that fungi contribute to many key ecosystem<br />

processes. This is particularly important in forest soils, where saprotrophic fungi<br />

are the main drivers <strong>of</strong> plant organic matter decomposition and ectomycorrhizal<br />

fungi facilitate nutrient uptake for the host plants, thereby affecting plant growth<br />

and fitness. Additionally, the presence <strong>of</strong> soil fungi shape seedling establishment<br />

and evidence suggests that ectomycorrhizal fungi may contribute to the distribution<br />

<strong>of</strong> photosynthate from one tree to another, directly regulating the survival <strong>of</strong><br />

nurse seedlings. Despite their important ecological roles, there is a paucity <strong>of</strong> information<br />

regarding taxonomic diversity and abundance associated with plant<br />

host. Largely this is due to the fact that many fungi are unculturable, lack known<br />

sexual structures, and are recorded only by environmental sequencing. In order to<br />

understand the diversity <strong>of</strong> fungi associated with differing forest types, we have<br />

utilized next-generation sequencing techniques (454/Roche and Illumina/Solexa)<br />

to probe soil fungal diversity. Seven monoculture forest types (Fagus, Quercus,<br />

Pinus, Pseudotsuga, Picea, an “old-growth” native forest dominated by Fagus,<br />

and an early successional Populus bio-energy plantation) were sampled. Nextgeneration<br />

sequencing showed extreme diversity <strong>of</strong> fungal taxa. Only 24% <strong>of</strong> taxonomic<br />

sequence reads were able to be identified with existing fungal databases.<br />

This finding agrees with recent studies employing new sequencing techniques to<br />

probe fungal diversity. Our sequencing shows an average <strong>of</strong> more than 2000<br />

unique fungal taxa in as little as four grams <strong>of</strong> forest soil. This presentation will<br />

discuss diversity across forest type surveyed and across seasonal sampling.<br />

Herrera, Cesar S 1 , Amy Y Rossman 2 , Gary J Samuels 2 , and Priscila Chaverri<br />

1 . 1 University <strong>of</strong> Maryland, Department <strong>of</strong> Plant Sciences and Landscape Architecture,<br />

2102 Plant Sciences Building, College Park, MD 20742, 2 United<br />

States Department <strong>of</strong> Agriculture, Systematic Mycology and Microbiology Lab,<br />

Rm. 246, Bldg. 010A, BARC-West, 10300 Baltimore Ave., Beltsville, MD<br />

20705. A new monotypic genus to accommodate Cosmospora vilior and related<br />

species<br />

Cosmospora sensu Rossman (Ascomycota, Hypocreales, Nectriaceae)<br />

has included nectroid fungi with small, reddish, KOH+, smooth, thin-walled, laterally<br />

collapsing when dry, non- or weakly stromatic perithecia. Recently, the<br />

group was found to be polyphyletic based on molecular data, and has been segregated<br />

into multiple genera. However, not all Cosmospora-like fungi have been<br />

treated systematically. Some <strong>of</strong> these species form a clade that includes C. vilior<br />

(= Nectria vilior) and many species <strong>of</strong>ten labeled as “Cosmospora sp.” The objectives<br />

<strong>of</strong> this research were to designate an epitype for N. vilior, which has been<br />

misapplied based on examination <strong>of</strong> the type specimen, and determine its phylogenetic<br />

position within Cosmospora sensu lato and the Nectriaceae. A multilocus<br />

phylogeny was constructed based on six loci (ITS, LSU, RPB1, MCM7, TEF1,<br />

and TUB) to estimate the species tree. Results from the phylogenetic analyses indicated<br />

that C. vilior forms a monophyletic group with other Cosmopora-like<br />

fungi that have an Acremonium-like anamorph and that parasitize Eutypa and Eutypella<br />

(Diatrypaceae). The group is phylogenetically distinct from other previously<br />

segregated genera, and for that reason, a new genus is described to accommodate<br />

these species.<br />

Herrera, Jose and Ravin Poudel. Department <strong>of</strong> Biology, 100 E. Normal, Truman<br />

State University, Kirksville, MO <strong>63</strong>501. The effect <strong>of</strong> soil on the fungal<br />

root endophyte community <strong>of</strong> Zea mays<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 21


Zea mays is, arguably, the most economically important grass domesticated<br />

by man. Yet, few studies have examined and fully characterized the fungal<br />

root endophyte community inhabiting the belowground portion <strong>of</strong> the plant. Taxonomic<br />

molecular identification <strong>of</strong> fungal species based upon the internal transcribed<br />

spacer (ITS) region <strong>of</strong> rDNA obtained from roots reveal soil as a prominent<br />

factor determining the composition <strong>of</strong> micr<strong>of</strong>ungal communities within the<br />

roots <strong>of</strong> maize (Zea mays subsp. mays, previously fully sequenced variety B73)<br />

and its progenitor, teosinte (Zea mays subsp. parviglumis), grown in Missouri clay<br />

soil or Missouri clay soil mixed with desert soil from New Mexico. Fungal communities<br />

in maize and teosinte grown in mixed soils were dominated by different<br />

proportions <strong>of</strong> Hypocreales spp. and dark septate endophytes (Pleosporales spp.,<br />

and Sordariales spp., primarily) while those grown only in Missouri clay soils<br />

were dominated by Glomerales spp. Moreover, fungal communities colonizing<br />

roots <strong>of</strong> Z. mays differed significantly from those colonizing roots <strong>of</strong> native (and<br />

co-habiting) fescue grass (Festuca arundinacea). Interestingly, Pleosporean sequences<br />

also include Paraphaeosphaeria spp., one <strong>of</strong> the dominant and common<br />

fungal species described in various grass species across North <strong>America</strong>. The presence<br />

<strong>of</strong> such cosmopolitan fungal root endophytes suggests that some fungal associations<br />

are general in nature and could function as evolutionary hinges that<br />

help plants develop adaptations to a variety <strong>of</strong> environmental challenges.<br />

Hess, Jaqueline 1 , Inger Skrede 2 , and Anne Pringle 1 . 1 Department <strong>of</strong> Organismic<br />

and Evolutionary Biology, Harvard University, 16 Divinity Ave., Cambridge,<br />

MA 02130, 2 Microbial Evolution Research Group (MERG), Department <strong>of</strong> Biology,<br />

University <strong>of</strong> Oslo, P.O. Box 1066 Blindern, N-0316 Oslo, Norway.<br />

Transposable element dynamics in ectomycorrhizal genomes: a close-up<br />

perspective<br />

Ectomycorrhizal (ECM) symbiosis between plants and fungi has evolved<br />

several times in the Fungal Kingdom. Ongoing genome sequencing <strong>of</strong> ectomycorrhizal<br />

(ECM) fungi is increasing the understanding <strong>of</strong> the genetic mechanisms<br />

central to the evolution <strong>of</strong> the symbiosis. Besides changes in gene content and<br />

gene family sizes, differences in genome architecture are also apparent, likely<br />

caused by changes in the ecology and population dynamics <strong>of</strong> ECM fungi. A<br />

common feature <strong>of</strong> ECM genomes sequenced to date is greatly elevated transposable<br />

element (TE) content: approximately 20% <strong>of</strong> the genome <strong>of</strong> Laccaria bicolor<br />

and 60% <strong>of</strong> the genome <strong>of</strong> Tuber melamsporum is made up <strong>of</strong> TEs. The<br />

genus Amanita encompasses an evolutionary transition from a free-living, saprotrophic<br />

to ECM niche and provides a powerful model for interrogating genome<br />

evolution across closely related, ecologically distinct species. To investigate<br />

changes in TE dynamics after evolution <strong>of</strong> the ECM niche we used next generation<br />

sequencing technologies to shotgun sequence the genomes <strong>of</strong> five species,<br />

covering the major clades <strong>of</strong> the genus Amanita (three ECMs and two saprotrophs)<br />

and the saprotrophic outgroup species Volvariella volvacea. We found the<br />

ECM genome sizes to be expanded, ranging from one and a half to nearly double<br />

the size compared to the saprotrophic genomes. Here, we will present our informatics<br />

approach for TE detection from short read data and fragmented assemblies<br />

and discuss the patterns <strong>of</strong> TE invasion in the different ECM clades <strong>of</strong> Amanita.<br />

Hicks, Sarah L 1 , Emily Farrer 2 , Robert Sinsabaugh 1 , Andrea Porras-Alfaro 1,3 ,<br />

and Katherine N Suding 2 . 1 Biology Dept, University <strong>of</strong> New Mexico, MSC03<br />

2020 1 University <strong>of</strong> New Mexico Albuquerque, NM 87105, 2 Environmental<br />

Science, Policy & Management University <strong>of</strong> California at Berkeley Berkeley,<br />

CA 94720, 3 Department <strong>of</strong> Biological Sciences Western Illinois University 1<br />

University Circle Waggoner Hall 372 Macomb, IL 61455. Effect <strong>of</strong> nitrogen<br />

pollution on root fungal endophyte communities from two co-dominant<br />

alpine tundra plants at Niwot, LTER<br />

Due to pollution from Boulder and Denver, CO, Nitrogen (N) deposition<br />

is increasing at Niwot Ridge, LTER. Over the last decade, one <strong>of</strong> the co-dominant<br />

alpine tundra plants, Geum rossii, has experienced substantial dieback, while the<br />

other co-dominant, Deschampsia caespitosa, gains dominance. G. rossii dieback<br />

has been attributed to elevated soil N rather than inter-species competition. The<br />

mechanism by which N kills G. rossii is still unknown. We hypothesize elevated<br />

soil N causes shifts in endophytic fungal communities, and that these shifts are responsible<br />

for G. rossii death. We expect G. rossii communities to become more<br />

pathogenic and parasitic in response to N. G. rossii and D. caespitosa roots were<br />

sampled from N-addition and control plots. Samples were surface sterilized prior<br />

to DNA extraction and 454 titanium pyrosequencing <strong>of</strong> fungal ITS ribosomal<br />

DNA. Sequences were clustered into OTUs by 97% similarity in Qiime. OTU<br />

correlation with N was assessed with ANOVA and community composition correlation<br />

with N was assessed with RDA, using the nlme and vegan packages in R<br />

respectively. G. rossii endophyte communities only clustered by N treatment at<br />

species level. Community shifts occurred among closely related species, and most<br />

sensitive taxa responded to N positively, including taxa that belong to pathogenic<br />

groups. Deschampsia caespitosa communities clustered by N at all taxonomic<br />

levels, and community shifts occurred between distantly related species, so that<br />

whole fungal Classes responded to N. Most affected taxa responded negatively.<br />

22 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

These data indicate that endophyte communities <strong>of</strong> D. caespitosa and G. rossii respond<br />

substantially differently to N addition. Deschampsia caespitosa endophyte<br />

communities appear to change more fundamentally. These changes may be important<br />

to D. caespitosa’s ability to adapt to changing nutrients. If so, these data<br />

may provide insight into the mechanisms by which abiotic soil factors alter above<br />

ground vegetation dynamics.<br />

Hobbie, Erik A. Earth Systems Research Center, University <strong>of</strong> New Hampshire,<br />

Durham, New Hampshire, USA. Assessing fungivory in rodents, marsupials,<br />

and kiwi from stable isotope evidence<br />

Fungivory is both a common dispersal mechanism for many fungi and an<br />

important source <strong>of</strong> nutrients and energy for many small animals, particularly rodents<br />

and marsupials. Studies <strong>of</strong> fungivory generally rely on scat or gut content<br />

analyses. An alternate technique, stable isotope measurements on animal protein,<br />

relies on sporocarps <strong>of</strong> ectomycorrhizal fungi having distinctive carbon and nitrogen<br />

isotope ratios ( 13 C/ 12 C and 15 N/ 14 N, expressed as d 13 C and d 15 N) compared<br />

to other ecosystem components. This approach allows retrospective studies<br />

by using archived museum specimens; here, I present three case studies <strong>of</strong> this approach.<br />

In studies <strong>of</strong> western Oregon rodents, mean isotopic values for the five<br />

sampled taxa separate into three main groups, with creeping vole and red tree vole<br />

quite similar (low in d 13 C and d 15 N, herbivorous), red-backed vole and flying<br />

squirrel similar (high in d13C and d15N, indicating high levels <strong>of</strong> fungivory), and<br />

Douglas squirrel different from the other two groupings (high in d 13 C, low in<br />

d 15 N, probably primarily a seed-eater). In ten genera <strong>of</strong> Australian marsupials,<br />

potoroos and bettongs appeared highly fungivorous, as expected from prior dietary<br />

studies. Kiwi in New Zealand may perform the same dispersal function for<br />

truffles as small mammals do elsewhere. Kiwis were isotopically distinct from<br />

frugivorous and insectivorous birds, but truffle consumption probably cannot be<br />

distringuished isotopically from consumption <strong>of</strong> soil invertebrates. Isotopic analysis<br />

appears to be a potentially useful tool to determine levels <strong>of</strong> fungivory in historic<br />

and current animal populations.<br />

Hodgins-Davis, Andrea 1 , Aleksandra Adomas 2 , Daniel Rice 3 , and Jeffrey<br />

Townsend 1 . 1 Department <strong>of</strong> Ecology and Evolutionary Biology, Yale University,<br />

New Haven, CT 06520, 2 National Institute <strong>of</strong> Environmental Health Sciences,<br />

Research Triangle Park, NC 27709, 3 Department <strong>of</strong> Organismic and Evolutionary<br />

Biology, Harvard University, Cambridge, MA 02138. Modelling transcriptional<br />

reaction norm evolution in the model yeast Saccharomyces cerevisiae<br />

Genetic variation for plastic phenotypes has the potential to contribute a<br />

large proportion <strong>of</strong> the phenotypic variation available to selection for novel adaptation.<br />

Models for linking population variation with plasticity make predictions<br />

about how selection may constrain the evolution <strong>of</strong> gene expression variation<br />

across transcriptional reaction norms. To empirically test these models, we have<br />

characterized population variation for transcriptional reaction norms in Saccharomyces<br />

cerevisiae for an ecologically relevant gradient <strong>of</strong> copper concentrations<br />

from starvation to toxicity. We find that although the vast majority <strong>of</strong> transcriptional<br />

variation is small in magnitude, not just some, but most genes demonstrate<br />

variable expression across environments or genetic backgrounds. Functionally,<br />

the most highly expressed genes defined three distinct cellular states across the<br />

copper reaction norm consistent with previous results characterizing eukaryotic<br />

responses to copper starvation, copper-replete fermentation, and copper overdose.<br />

These cellular states included direct transcriptional responses to intracellular levels<br />

<strong>of</strong> copper ions and diverse indirect metabolic adjustments to the consequences<br />

<strong>of</strong> changed copper levels. Indirect homeostatic changes <strong>of</strong> expression were more<br />

variable among genotypes in their direction <strong>of</strong> response than were the reaction<br />

norms <strong>of</strong> genes directly regulated by copper-binding transcription factors. To interpret<br />

this variability in transcriptional reaction norms in the context <strong>of</strong> the<br />

processes <strong>of</strong> natural selection and neutral drift, we account for variation in mutation<br />

and regulatory degree across the genome by parameterizing classic models <strong>of</strong><br />

phenotypic evolution. Empirically estimating population genetic variance, mutational<br />

variance, and regulatory degree, we infer the strength <strong>of</strong> stabilizing selection<br />

operating on gene expression levels. We present evidence consistent with either<br />

weak or infrequent stabilizing selection on most gene expression phenotypes,<br />

and discuss the implications <strong>of</strong> the lack <strong>of</strong> constraint for the evolvability <strong>of</strong> transcriptional<br />

reaction norms.<br />

Hodkinson, Brendan P 1 and James C Lendemer 2 . 1 International Plant Science<br />

Center, The New York Botanical Garden, 2900 Southern Blvd., Bronx, NY<br />

10458-5126, 2 Institute <strong>of</strong> Systematic Botany, The New York Botanical Garden,<br />

2900 Southern Blvd., Bronx, NY 10458-5126. A modern, high-throughput<br />

workflow for biodiversity research integrating floristics, taxonomy, phylogeny,<br />

ecology and conservation<br />

Continued on following page


The field <strong>of</strong> Systematic Biology has increasingly become fragmented in<br />

recent years, with many scientists focusing in on a small number <strong>of</strong> methods or<br />

approaches for studying biodiversity and its origins. Here we present an integrated<br />

workflow for studying the many facets <strong>of</strong> biodiversity and systematics. This<br />

workflow is currently being used to conduct a large-scale inventory <strong>of</strong> lichens in<br />

the Mid-Atlantic Coastal Plain <strong>of</strong> the United States that includes the collection and<br />

analysis <strong>of</strong> ~35,000 new specimens. Using floristic habitat sampling (FHS), specimens<br />

representing the organismal group <strong>of</strong> interest are collected from individual<br />

sites throughout the study region, and are subsequently analyzed anatomically and<br />

chemically. Specimens that cannot be identified taxonomically with anatomical/chemical<br />

analyses alone are sequenced using a multi-locus 454-based approach.<br />

Sequences from the genes <strong>of</strong> interest (in our case, mtSSU, nucLSU and<br />

nucITS) are compared with publicly available reference sequences to determine<br />

the phylogenetic affinities <strong>of</strong> species that remain either unnamed or ‘incertae<br />

sedis.’ All collections are databased rapidly using modular databasing s<strong>of</strong>tware<br />

(KE EMu). Large-scale ecological analyses are performed on database-generated<br />

taxon lists by hijacking data management s<strong>of</strong>tware originally designed for organizing<br />

and analyzing large molecular sequence data sets. Overall results are compiled<br />

and made available on the web in the form <strong>of</strong>, e.g., checklists (for regions,<br />

sub-regions and sites), detailed taxonomic treatments, molecular sequence alignments,<br />

phylogenetic trees, and specimen-based ecological data sets. Both through<br />

these online resources and workshops/forays for the public, information is disseminated<br />

for integration into conservation and management plans to preserve<br />

biodiversity for the future.<br />

Hoppe, Björn 1 , Tiemo Kahl 2 , François Buscot 1 , and Dirk Krüger 1 . 1 UFZ -<br />

Helmholtz Centre for Environmental Research, Department <strong>of</strong> Soil Ecology, Th.-<br />

Lieser-Str. 4, D-06120 Halle/Saale, Germany, 2 University <strong>of</strong> Freiburg, Institute<br />

<strong>of</strong> Silviculture, Tennenbacherstr. 4, D-79085 Freiburg i. Brsg., Germany. A study<br />

<strong>of</strong> correlations between fungal and bacterial communities in dead wood<br />

Dead wood (a.k.a. coarse woody debris, CWD) is a major habitat in forest<br />

ecosystems that is home and food for various types <strong>of</strong> organisms. Usually, researchers<br />

study the fungi and xylobiontic insects present and active in the wood<br />

decay process. The role <strong>of</strong> bacteria in the decomposition <strong>of</strong> this complex substrate<br />

is yet underexplored. One aspect <strong>of</strong> the project FunWood <strong>of</strong> the DFG (German<br />

Science Foundation) Biodiversity Exploratories is to investigate interrelations between<br />

fungal and bacterial diversity on CWD <strong>of</strong> European beech, Norway spruce<br />

and Scots pine along a gradient <strong>of</strong> forest sites that were under different silvicultural<br />

management practices. The same DNA extractions from drilled CWD<br />

served as templates for PCR subsequently processed for functional and diversity<br />

fingerprints (nitrogenase nifH T-RFLP, ARISA) as well as Sanger sequencing.<br />

Preliminary results entail 16S rDNA sequences from highly diverse eubacterial<br />

communities in CWD comprising 10 different classes. The detected families Beijerinckiaceae,<br />

Rhizobiaceae and Bradyrhizobiaceae are known to include potent<br />

nitrogen fixing bacteria. Conceivably, the extreme C/N ratios <strong>of</strong> CWD reaching<br />

350-800:1 make it increasingly harder for wood decay fungi to meet their nitrogen<br />

demands as the wood decay process goes on. Thus, we aim to test further for<br />

the presence <strong>of</strong> actually nitrogen fixing bacteria and their potential role in filling<br />

the fungal nitrogen needs in this habitat. We hitherto found 26 different nifH sequences<br />

from 5 logs. None was closer than 97% similarity to any on GenBank,<br />

indicating the presence <strong>of</strong> uncharacterized, potentially unknown bacteria. By the<br />

time <strong>of</strong> presentation, statistically analyzed nifH sequence data <strong>of</strong> 48 CWD logs are<br />

ready to be discussed, along with fungal diversity data based on sporocarps and<br />

DNA.<br />

Horn, Bruce W 1 , Ronald B Sorensen 1 , Marshall C Lamb 1 , Victor S Sobolev 1 ,<br />

and Ignazio Carbone 2 . 1 National Peanut Research Laboratory, USDA, ARS,<br />

Dawson, GA, 39842, 2 Department <strong>of</strong> Plant Pathology, North Carolina State University,<br />

Raleigh, NC, 27606. Formation <strong>of</strong> Aspergillus flavus sclerotia on corn<br />

grown under different drought stress conditions<br />

Aspergillus flavus is a major producer <strong>of</strong> carcinogenic aflatoxins worldwide<br />

in corn, peanuts, tree nuts, cottonseed, spices and other crops. Many countries<br />

have strict limits on the amount <strong>of</strong> aflatoxins permitted in human commodities<br />

and animal feed. Sclerotia produced by A. flavus serve several functions in the<br />

life cycle <strong>of</strong> the fungus: (1) resistance to adverse environmental conditions; (2)<br />

colonization <strong>of</strong> substrates through myceliogenic and sporogenic germination; and<br />

(3) sexual reproduction. Drought stress in corn results in an increase in aflatoxin<br />

contamination, but little is known about the effects <strong>of</strong> drought on sclerotium production.<br />

Corn was grown at Shellman, Georgia, in 2010 under different drought<br />

stress conditions consisting <strong>of</strong>: rainfall only (0% irrigation) and 33, 66 and 100%<br />

<strong>of</strong> the recommended irrigation amounts. Decreasing water availability showed a<br />

progressive reduction in grain yield (r = 0.96, p < 0.001) and an increase in aflatoxin<br />

concentration (r = –0.62, p < 0.05), but had no significant effect on percentage<br />

<strong>of</strong> ears showing Aspergillus section Flavi sporulation or percentage <strong>of</strong> ears<br />

with sclerotia. Of the 8706 ears examined, sclerotia (n = 6022) were recovered<br />

from 84 ears (1%). Sclerotia <strong>of</strong> A. flavus L (large sclerotial) strain were dominant<br />

(78 ears); less common were sclerotia <strong>of</strong> A. flavus S (small sclerotial) strain (8<br />

ears), A. parasiticus (3 ears) and A. alliaceus (2 ears). The sexual stage was not<br />

detected in any sclerotia, including those <strong>of</strong> homothallic A. alliaceus, suggesting<br />

that sclerotia may require additional incubation after dispersal onto soil.<br />

Horton, Thomas R 1 , Martin A Nuñez 2 , Yazmin Rivera 1 , G Amico 2 , R Dimarco<br />

2 , N Barrios 2 , Jeremy Hayward 1 , and Dan Simberl<strong>of</strong>f 2 . 1 Department <strong>of</strong> Environmental<br />

and Foresty Biology, SUNY-ESF, Syracuse, NY 13210, 2 Department<br />

<strong>of</strong> Ecology and Evolutionary Biology, University <strong>of</strong> Tennessee, Knoxville, TN<br />

37919. On the role <strong>of</strong> ectomycorrhizal fungi in Pinaceae invasions<br />

Establishment <strong>of</strong> Pinaceae in locations with no history <strong>of</strong> the trees requires<br />

the co-introduction <strong>of</strong> compatible ectomycorrhizal (EM) fungi, which are<br />

also required to facilitate invasions. Compatible fungi disperse from plantations<br />

through mycelial networks that grow from plantation edges (short distance), and<br />

spores dispersed by wind, water and mammals (potentially long distance). Invasion<br />

<strong>of</strong> Pinaceae from plantations into native forests has been limited on Puerto<br />

Rico (USA) and Isla Victoria (Argentina). The trees do not associate with native<br />

EM fungi to any great extent, if at all. On Puerto Rico most established pines are<br />

within decimeters <strong>of</strong> plantations, but some trees have established at 1000m. Root<br />

tips collected from established pines at 1000m from plantations yielded Rhizopogon<br />

spp., as well as Pisolithus tinctorius, Wilcoxina mikolae and a member <strong>of</strong><br />

the Atheliaceae. This indicates mammalian (Rhizopogon) as well as other dispersal<br />

vectors (Pisolithus, Wilcoxina, Atheliaceae). Bioassays with dried soils collected<br />

in areas 1000m from plantations and without pines only yielded Rhizopogon.<br />

Dispersal <strong>of</strong> fungi may be facilitated by rats and mice introduced since the<br />

1500s. In addition, tropical storms and anthropogenic activity may vector the<br />

fungi over long distances. On Isla Victoria establishment is largely restricted<br />

around plantations. We have not observed the rare native deer or any rodents in<br />

or near plantations. However, nonnative deer were first introduced with conifers<br />

about 100 years ago and pigs have become numerous since 2004. A bioassay with<br />

ponderosa pine and Douglas fir using fecal pellets from the nonnative mammals<br />

collected away from plantations yielded Rhizopogon, Suillus, and Hebeloma mycorrhizae.<br />

Severe weather during the peak fruiting period is probably less important<br />

for spore dispersal on Isla Victoria than it is on Puerto Rico. We hypothesize<br />

the degree long-distance dispersal and accumulation <strong>of</strong> compatible propagules<br />

can impact the rate at which these invasions proceed.<br />

Hughes, Karen W and Ronald H Petersen. Ecology and Evolutionary Biology,<br />

University <strong>of</strong> Tennessee, Knoxville, TN 37996. Current fungal data from the<br />

Great Smoky Mountains ATBI are inadequate to determine conservation<br />

status: what is needed?<br />

Rapid urbanization and deforestation in the Southeastern US has heavily<br />

impacted native plants, and many remain only in small reserves. While there is no<br />

similar assessment for fungi, it is likely that fungal species follow similar patterns<br />

<strong>of</strong> fragmentation and localized extirpation. Determining endemism and rarity,<br />

however, is significantly hampered by a lack <strong>of</strong> focused species-level research<br />

and data concerning fungal distribution patterns. The Great Smoky Mountains<br />

National Park (GSMNP) is a protected reserve that exhibits high biological diversity<br />

for fungi, salamanders and plants. The GSMNP and adjacent regions <strong>of</strong><br />

the southern Appalachian Mountains provide an opportunity to identify remnants<br />

<strong>of</strong> native southeastern mycota. We report studies at the species level that identify<br />

putative endemics for further study, examine distributions in so far as is possible<br />

and note high levels <strong>of</strong> cryptic speciation in fungi that may influence estimates <strong>of</strong><br />

richness and endemism.<br />

Hung, Richard. Plant Biology, Rutgers University, 59 Dudley Rd., New<br />

Brunswick, NJ 08901. Trichoderma viride volatile organic compound pr<strong>of</strong>ile:<br />

The effect <strong>of</strong> individual components on Arabidopsis thaliana<br />

Several Trichoderma species are successful biocontrol agents. Their effects<br />

include plant growth promotion, stimulation <strong>of</strong> systemic defense responses,<br />

and protection from pathogens. Our research has shown that the volatile organic<br />

compounds (VOCs) <strong>of</strong> Trichoderma viride have plant growth promoting effects<br />

on Arabidopsis thaliana without physical contact. GC-MS analysis <strong>of</strong> the VOC<br />

pr<strong>of</strong>ile <strong>of</strong> T. viride has resulted in 51 unique detectable compounds <strong>of</strong> which<br />

isobutyl alcohol, isopentyl alcohol, and 3-methylbutanal were the most abundant.<br />

Experiments have been performed to determine which components <strong>of</strong> the T. viride<br />

pr<strong>of</strong>ile are responsible for the observed growth in A. thaliana. These compounds,<br />

in addition to mushroom alcohol (1-octen-3-ol) and geosmin, have been<br />

tested on model systems with A. thaliana. Exposure to 1-octen-3-ol has resulted<br />

in increased growth both above and below ground. The average root mass <strong>of</strong> control<br />

plants was 0.36g and the average mass <strong>of</strong> VOC exposed plants was 0.77g<br />

showing a 113% increase in plant mass. In addition there was a 60% increase in<br />

chlorophyll concentration (5.5mg/g control, 8.8mg/g test). Conductivity readings<br />

<strong>of</strong> the tested plants were the same as controls at 12% ± 2% indicating no cell wall<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 23


instability. Trypan blue staining showed no necrotic lesions and hydrogen peroxide,<br />

a reactive oxygen species (ROS) that is precursor to necrotic lesions was not<br />

detected with 3,3’-diaminobenzidine (DAB) staining. Exposure to ethanol, another<br />

component <strong>of</strong> the VOC pr<strong>of</strong>ile, resulted in reduced seedling formation<br />

(73%) as compared to control (92%). More seeds were unable to throw <strong>of</strong>f their<br />

seed coat and complete seedling formation (15%) than control (2%) at the end <strong>of</strong><br />

three days.<br />

Hustad, Vincent P 1,2 and Andrew N Miller 2 . 1 Dept. <strong>of</strong> Plant Biology, University<br />

<strong>of</strong> Illinois, 505 South Goodwin Ave., Urbana, IL 61801, 2 Illinois Natural History<br />

Survey, University <strong>of</strong> Illinois, 1816 South Oak St., Champaign, IL 61820.<br />

Studies in Geoglossomycetes systematics<br />

Geoglossomycetes are a widespread and diverse class <strong>of</strong> fungi known<br />

from every continent except Antarctica. Commonly referred to as earth-tongues<br />

due to their morphology and terrestrial habitat, these fungi have long been a subject<br />

<strong>of</strong> mycological interest, although few molecular systematic studies have been<br />

conducted within the group. This research presents the latest results <strong>of</strong> our attempt<br />

to generate the first modern, multi-gene, species-level phylogeny within Geoglossomycetes.<br />

Sequences <strong>of</strong> the partial nuclear ribosomal 28S large subunit and<br />

internal transcribed spacer genes, as well as the protein-coding genes, RPB1 and<br />

MCM7, were generated for numerous species previously thought to belong in Geoglossomycetes.<br />

Results <strong>of</strong> maximum parsimony, maximum likelihood, and<br />

Bayesian analyses confirmed previous findings that Geoglossomycetes is a wellsupported<br />

monophyletic class that includes the genera Geoglossum, Nothomitra,<br />

Sarcoleotia, and Trichoglossum. Polyphyletic genera within Geoglossomycetes<br />

were also observed and two new genera, Glutinoglossum and Sabuloglossum,<br />

were established.<br />

Hutchinson, Miriam I 1 , Amy J Powell 2 , Kylea J Parchert 2 , Joanna L Redfern 1 ,<br />

Andrea M Martinez 1 , Randy M Berka 3 , Eric Ackerman 2 , Blake Simmons 4,5 ,<br />

Igor V Grigoriev 6 , and Donald O Natvig 1 . 1 Department <strong>of</strong> Biology, University<br />

<strong>of</strong> New Mexico, Albuquerque, NM 87131, 2 Sandia National Laboratories, Albuquerque,<br />

NM 87123, 3 Novozymes, Inc., Davis, CA 95618, 4 Sandia National<br />

Laboratories, Livermore, CA 94551, 5 Joint BioEnergy Institute, Emeryville, CA<br />

94608, 6 DOE Joint Genome Institute, Walnut Creek, CA 94598. Reproductive<br />

genetics and development in the fungus Myceliophthora heterothallica, a<br />

thermophilic model for the Chaetomiaceae<br />

Members <strong>of</strong> the Chaetomiaceae are among the most reported fungi in<br />

studies <strong>of</strong> biomass degradation. They are <strong>of</strong> interest for their abilities to produce<br />

thermostable carbohydrate-active enzymes, which led to the sequencing <strong>of</strong><br />

genomes from two thermophilic species, Myceliophthora thermophila and<br />

Thielavia terrestris. Until now, there has been no genetically tractable model either<br />

for this family, or more generally, for thermophilic fungi. We have characterized<br />

reproduction in the thermophile Myceliophthora heterothallica towards<br />

the goals <strong>of</strong> establishing this organism as a model for the group and developing it<br />

as an expression platform. M. heterothallica was reported to be heterothallic<br />

based on the fact that matings between two strains resulted in the production <strong>of</strong><br />

fruiting bodies and ascospores. Prior to our work, however, heterothallism had not<br />

been confirmed with independent assortment <strong>of</strong> mating loci and autosomal genes.<br />

We speculate that this lack <strong>of</strong> confirmation <strong>of</strong> true heterothallism resulted from a<br />

failure to obtain ascospore germination. We found that ascospores are resistant to<br />

germination at temperatures below 47-50 ° C. This discovery allowed us to confirm<br />

heterothallism and analyze the segregation <strong>of</strong> markers in crosses. Sequences<br />

from opposite mating types show that mating regions are conserved relative to<br />

other Sordariales. Interestingly, different stages <strong>of</strong> development have different<br />

temperature optima: ascospore germination occurs at 47 ° C and above, ascocarp<br />

formation is optimal at 30 ° C, and growth is optimal at 45 ° C. We have successfully<br />

crossed M. heterothallica strains from Indiana, New Mexico and Germany,<br />

and we are expanding the number <strong>of</strong> known strains by surveying across latitudinal<br />

and elevation gradients. In addition, we are developing methods for transformation<br />

and gene replacement. Our goal is to develop M. heterothallica as a model<br />

organism to study fundamental aspects <strong>of</strong> thermophily and the biology <strong>of</strong> Chaetomiaceae.<br />

Hwang, Jonathan 1 , Qi Zhao 2 , Zhuliang Yang 2 , Zheng Wang 1 , and Jeffrey P<br />

Townsend 1 . 1 Ecology and Evolutionary Biology Department, Yale University,<br />

165 Prospect Street, New Haven, CT 06520, USA, 2 Key Laboratory <strong>of</strong> Biodiversity<br />

and Biogeography, Kunming Institute <strong>of</strong> Botany, Chinese Academy <strong>of</strong><br />

Sciences, Kunming, Yunnan, China. Ecological diversification <strong>of</strong> saddle fungi:<br />

Helvella indicated by phylogeny and secondary structure <strong>of</strong> ITS sequences<br />

from annotated collections and environmental samples<br />

Saddle fungi Helvella (Helvellaceae, Pezizomycotina) species are morphologically<br />

well-defined and have been intensively studied for their taxonomy<br />

and classification. Ecological roles as being saprotrophic or mycorrhizal have<br />

been suggested with indirect evidence, including molecular data from environmental<br />

samples, for some species in the genus. Recent molecular phylogenies<br />

24 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

based on internal transcribed spacer <strong>of</strong> rDNA (ITS) did not support morphological<br />

groupings; however, a robust phylogeny <strong>of</strong> ITS was not achieved due the large<br />

divergence observed among Helvella ITS sequences, which makes a dependable<br />

alignment <strong>of</strong> representative ITS sequences in the genus out <strong>of</strong> reach <strong>of</strong> traditional<br />

alignment programs. In this study, we applied SATé, a s<strong>of</strong>tware that combines<br />

progressive alignment and tree building to build so far the most inclusive ITS<br />

alignment for annotated Helvella species and their related environmental ITS sequences.<br />

Estimated secondary structures for selected ITS2 sequences verified the<br />

robustness <strong>of</strong> the phylogeny based on the best SATé alignment. Our results suggested<br />

a high homoplasious similarity in key morphologies in Helvella, and morphological<br />

groupings <strong>of</strong> the genus do not reflect the true evolutionary histories<br />

within the group. ITS phylogeny including environmental samples also provides<br />

strong evidence that ectomycorrhizal association is maintained in a clade <strong>of</strong><br />

Helvella and diversified to adapt for different hosts and geographic conditions.<br />

However, saprotrophic life styles in other Helvella species requires expanded<br />

metagenomic investigation on plant materials and soil samples, as no single ITS<br />

sequence was recovered from studies examining various soil samples for those<br />

common fungi.<br />

Isiloglu, Mustafa, Hakan Alli, and Hayrunisa Bas Sermenli. Mugla University<br />

Dept. <strong>of</strong> Biology, Campus, Mugla, Turkey. The Ecology <strong>of</strong> Morchella anatolica<br />

Morchella anatolica Isiloglu, Spooner, Alli & Solak was described in<br />

2010 from a very interesting area in Mugla, Turkey. In this paper, its remarkable<br />

and unique ecology is discussed.<br />

James, Timothy Y 1* , Adrian Pelin 2 , Stefan Amyotte 2 , Courtney S Frye 1 , Nicolas<br />

Corradi 2 , and Jason E Stajich 3 . 1 Dept. <strong>of</strong> Ecology and Evol. Biol., University<br />

<strong>of</strong> Michigan, Ann Arbor, MI 48109 USA, 2 Dept. <strong>of</strong> Biology, University <strong>of</strong> Ottawa,<br />

Ottawa, ON, Canada K1N 6N5, 3 Dept. <strong>of</strong> Plant Pathology and<br />

Microbiology, University <strong>of</strong> California, Riverside, CA 92521 USA. The biology<br />

and phylogeny <strong>of</strong> Rozella (phylum Cryptomycota) revealed by genome sequencing<br />

The enigmatic genus Rozella is a chytrid-like endoparasite <strong>of</strong> water molds<br />

that is related to a diverse array <strong>of</strong> organisms (Cryptomycota) known only from<br />

environmental DNA sequences. Rozella and other Cryptomycota are suggested to<br />

lack a chitinous cell wall during at least part <strong>of</strong> their trophic phase. We produced<br />

a draft genome sequence for Rozella using a combination <strong>of</strong> next generation sequencing<br />

technologies. Rozella possesses at least 5 chitin synthase genes that it<br />

uses during both resting sporangium formation and host invasion. The mitochondrial<br />

genome <strong>of</strong> Rozella is reduced in size and gene content, displaying a genome<br />

remarkably convergent with that <strong>of</strong> Plasmodium. A phylogenomic analysis confirms<br />

the placement <strong>of</strong> Rozella on the primary (earliest diverging) branch <strong>of</strong> the<br />

fungal kingdom along with other Cryptomycota such as microsporidia. The nuclear<br />

genome, however, has not undergone the major reduction in coding potential<br />

as observed in microsporidia.<br />

Jarvis, Susan G 1,2 , Andy F S Taylor 1,2 , Ian J Alexander 1 , and Steve Woodward<br />

1 . 1 School <strong>of</strong> Biological Sciences, University <strong>of</strong> Aberdeen, Aberdeen, AB24<br />

3UU, UK, 2 The James Hutton Institute, Craigiebuckler, Aberdeen, AB15 8QH,<br />

UK. Community composition <strong>of</strong> ectomycorrhizal fungi along a climatic gradient<br />

in Scotland<br />

Ectomycorrhizal (ECM) fungi play fundamental roles in forest ecosystems<br />

through tree nutrition and nutrient cycling. Understanding the responses <strong>of</strong><br />

ECM fungi to environmental change will be important to maintain forest productivity<br />

for the future. Assessing the distribution <strong>of</strong> ECM communities along existing<br />

large-scale environmental gradients provides an insight into the environmental<br />

factors driving community composition. Previous studies have suggested<br />

nitrogen deposition to be the main environmental driver <strong>of</strong> ECM community variation<br />

over large-scale gradients in Europe. However, these studies have included<br />

areas with very high nitrogen deposition, which can dramatically alter ECM communities,<br />

and this may have masked any climatic effect on these communities. In<br />

Scotland nitrogen deposition varies with latitude but is less severe than in continental<br />

Europe, with total nitrogen deposition below 20 kg N ha -1 yr -1 in most<br />

areas. Utilising the latitudinal gradient in nitrogen deposition and a longitudinal<br />

climate gradient a survey <strong>of</strong> the ECM fungi present on the roots <strong>of</strong> native Scots<br />

pine (Pinus sylvestris) across Scotland was conducted to assess the relative impacts<br />

<strong>of</strong> climate and moderate nitrogen deposition on ECM community composition.<br />

Ectomycorrhizal pine roots were collected from 15 native pinewoods and<br />

fungal species were identified from root tips by sequencing <strong>of</strong> the ITS rDNA region.<br />

Non-metric multidimensional scaling <strong>of</strong> community composition and vector<br />

fitting <strong>of</strong> 17 climatic, nitrogen and soil chemistry variables identified temperature<br />

and rainfall variables to be correlated with the primary axis <strong>of</strong> community<br />

variation. Nitrogen deposition was shown to have little influence on community<br />

Continued on following page


composition, although there was some effect <strong>of</strong> total soil nitrogen. This work<br />

highlights the potential for climate to influence ECM fungal communities and the<br />

need to understand the impacts, and possible interactions, <strong>of</strong> multiple drivers on<br />

community composition.<br />

Je, Hee Jeong, Chak Han Im, Na Hee Kwon, Mi Jin Kim, Min Keun Kim, Ki<br />

Kwan Park, and Jae San Ryu. Eco-friendliness Research Department,<br />

Gyeongsangnam-do Agricultural Research and Extension Services, Jinju 660-<br />

360, Republic <strong>of</strong> Korea. Colorimetric method for discriminating abnormal<br />

strains <strong>of</strong> Flammulina velutipes and SSR markers linked to the defective<br />

symptoms<br />

Winter mushroom, F. velutipes is one <strong>of</strong> the most popular mushrooms in<br />

Asian countries, including Korea, Japan, and China. The strain is preserved on<br />

slants or under mineral oil in mushroom farms, but occasionally it degenerates by<br />

unknown reasons, resulting in loss in yield and malformation. Therefore, a<br />

method is needed that is easy enough to be managed by farmers to detect abnormal<br />

strains using a pH indicator and carbon source. Media including MCM, PDA,<br />

YP supplemented with 0.5% <strong>of</strong> carbon sources (sawdust, lactose Avicel, CMC,<br />

etc) and pH indicators (bromothymol blue, congo red, methyl orange, bromocresol<br />

green, methyl red, bromocresol purple, phenol red, and bromophenol blue)<br />

were tested for addressing difference between normality and abnormality <strong>of</strong><br />

strains. YP supplemented popular sawdust and BTB was useful to detect 2 highly<br />

abnormal and 6 slightly abnormal strains that were undetectable by the previous<br />

method. 275 SSRs were identified to design primers and to develop markers<br />

linked to the defective symptoms. 2 SSR markers showing a polymorphism between<br />

normal and abnormal strains were screened. The difference in length <strong>of</strong><br />

bands was just 4 - 8 bases, so an automatic electrophoresis system will be required<br />

to detect such a small difference.<br />

Jenkinson, Thomas S 1,2 , Brian A Perry 3 , Rainier E Schaefer 1 , and Dennis E<br />

Desjardin 1 . 1 Department <strong>of</strong> Biology, San Francisco State University, 1600 Holloway<br />

Ave., San Francisco, CA 94132, 2 Department <strong>of</strong> Ecology and Evolutionary<br />

Biology, University <strong>of</strong> Michigan, 830 North Universty Ave., Ann Arbor, MI<br />

48109, 3 Department <strong>of</strong> Biology, University <strong>of</strong> Hawaii, Hilo, 200 West Kawili St.,<br />

Hilo, HI 96720. Cryptomarasmius gen. nov. is proposed in the Physalacriaceae<br />

to accommodate members <strong>of</strong> Marasmius section Hygrometrici<br />

Phylogenetic placement <strong>of</strong> the infrageneric section Hygrometrici (genus<br />

Marasmius sensu stricto) in prior molecular phylogenetic studies have been unresolved<br />

and problematical. Molecular analyses based on newly generated ribosomal<br />

nuc-LSU and 5.8S sequences resolve members <strong>of</strong> section Hygrometrici to<br />

the family Physalacriaceae, within the Agaricales. The new genus Cryptomarasmius<br />

is proposed in the Physalacriaceae to accommodate members <strong>of</strong> Marasmius<br />

section Hygrometrici. Taxonomic transfers <strong>of</strong> fourteen species within section Hygrometrici<br />

are recommended to reflect phylogenetic affinity.<br />

Johnson, Eric M, Eric D Tretter, Prasanna Kandel, and Merlin M White. Boise<br />

State University, Dept. <strong>of</strong> Biological Sciences, 1910 University Drive, Boise, ID<br />

83725. Molecular-based phylogenetic placement <strong>of</strong> the Asellariales, an enigmatic<br />

order <strong>of</strong> arthropod gut endosymbionts<br />

The order Asellariales is comprised <strong>of</strong> three genera, Asellaria, Orcheselleria,<br />

and Baltomyces, which are endosymbionts in the guts <strong>of</strong> isopods and springtails.<br />

Ultrastructural studies suggest placement <strong>of</strong> Asellariales in the subphylum<br />

Kickxellomycotina, and life history further suggests a relationship to the Harpellales,<br />

an order <strong>of</strong> insect gut endosymbionts. However, placement <strong>of</strong> the Asellariales<br />

at any level has eluded verification due to a lack <strong>of</strong> molecular data. Attempts<br />

to obtain gene sequences for any member <strong>of</strong> the Asellariales have been seriously<br />

hindered for a variety reasons, including sparse DNA samples due to the microscopic<br />

size <strong>of</strong> the specimens; lack <strong>of</strong> culturability; and contaminating DNA from<br />

the host animal as well as other fungi, bacteria, and even host food material in the<br />

gut. Here we report the successful amplification and sequencing <strong>of</strong> partial sequences<br />

<strong>of</strong> 5 genes (SSU and LSU rDNA, RPB1, RPB2, B-tubulin) for Asellaria<br />

ligiae, as well as a single gene (LSU rDNA) for Orchesellaria mauguoi. We<br />

also describe the development and use <strong>of</strong> custom primers to obtain the sequences,<br />

as well as the unusual nature <strong>of</strong> A. ligiae rDNA and challenges associated with it.<br />

Finally, we can place representatives <strong>of</strong> the Asellariales using Bayesian inference<br />

on the basis <strong>of</strong> molecular data from Asellaria and Orchesellaria.<br />

Johnson, Lynnaun 1 , Shiloh Lueschow 1 , Jason Stonewall 1 , Robert McCleery 2 ,<br />

Rod McClanahan 3 , Joseph Kath 4 , and Andrea Porras-Alfaro 1 . 1 Biological Sciences,<br />

Western Illinois University, Macomb, IL 61455, 2 Wildlife Ecology and<br />

Conservation PO Box 110430, University <strong>of</strong> Florida, Gainesville, FL 32611-<br />

0430, 3 4<br />

USDA Forest Service, Shawnee National Forest, Harrisburg, IL 62946,<br />

Illinois Department <strong>of</strong> Natural Resources, Springfield, IL 62702-1271. Description<br />

<strong>of</strong> mycobiomes associated with three species <strong>of</strong> bats in Illinois<br />

Geomyces destructans is the causative agent <strong>of</strong> Geomycosis better known<br />

as White Nose Syndrome (WNS). This fungus was documented in 2009 as a<br />

novel fungal species. Since 2006 it has killed more than a million bats to date with<br />

losses in bat populations <strong>of</strong> 75% or more in the northeastern United States.<br />

Though hypotheses suggest that G. destructans originated in Europe, the fungus<br />

currently shows no serious threat to European bats. With the high mortality<br />

caused by WNS in North <strong>America</strong>, much focus has been placed on the fungus G.<br />

destructans and current literature is lacking detailed analysis about the composition<br />

and structure <strong>of</strong> fungal communities (mycobiomes) in healthy bats. The main<br />

objectives <strong>of</strong> this project are 1) to document fungal flora <strong>of</strong> hibernating bats in<br />

northern IL caves, 2) to determine the presence <strong>of</strong> other potential Geomyces<br />

species and pathogens amongst bats. We sampled 30 bats including Myotis<br />

septentrionalis, Myotis sodalis, Perimyotis subflavus from four hibernacula within<br />

Illinois (before WNS arrival) and we isolated and sequenced 141 fungi using<br />

ITS primers. We found approximately 53 OTUs at 97% similarity. Fungi associated<br />

with bats were primarily dominated by Ascomycota (72% <strong>of</strong> the OTUs).<br />

Dominant orders include Eurotiales, Hypocreales, Capnodiales, Pleosporales and<br />

Sordariales. Common species isolated from bats included Cladosporium, Geomyces,<br />

Fusarium, Mortierella, Penicillium and Trichosporon. Thirteen Geomyces<br />

isolates were obtained from different bat species and phylogenetic analysis<br />

indicated at least three potential novel Geomyces clades. Temperature tests<br />

showed all Geomyces isolates were psychrotolerant, different than G. destructans,<br />

a psychrophilic fungus. The widespread distribution <strong>of</strong> Geomyces among bat mycobiomes<br />

can lead to false positives when detecting WNS. The role <strong>of</strong> these diverse<br />

bat mycobiome and their potential pathogenicity and interactions with G.<br />

destructans will require further analysis.<br />

Jumper, Katherine E 1 , Nishanta Rajakaruna 2 , and David Porter 3 . 1 122 Cottage<br />

Street Bar Harbor, ME 04609, 2 College <strong>of</strong> the Atlantic 105 Eden Street Bar Harbor,<br />

ME 04609, 3 Department <strong>of</strong> Plant Biology University <strong>of</strong> Georgia Athens, GA<br />

30602. Diversity and heavy metal content <strong>of</strong> macr<strong>of</strong>ungi found on serpentine<br />

andgranite outcrops, Deer Isles, Maine, USA<br />

Fungal diversity in relation to edaphic conditions has rarely been studied.<br />

We are conducting a year-long comparative survey <strong>of</strong> macr<strong>of</strong>ungi on serpentine<br />

and granite outcrops to determine whether there are differences in fungal biodiversity<br />

based on edaphic factors, and whether fungi tissue differ in heavy metal<br />

content based on the substrate upon which they are found. Previous studies we<br />

have conducted document differences in diversity <strong>of</strong> lichens, bryophytes, and vascular<br />

plants at these two sites, including significant differences in heavy metal accumulation<br />

in vascular plant species found at both sites. Serpentine soils contain<br />

high amounts <strong>of</strong> heavy metals and low amounts <strong>of</strong> essential nutrients and are a<br />

challenging environment for plant growth. Serpentine outcrops are known to harbor<br />

many rare and endemic plants, but very little is known about the ecology <strong>of</strong><br />

macr<strong>of</strong>ungi in serpentine habitats. We predict that we will find differences in<br />

species diversity between the two outcrops, reflecting differing abilities in the fungal<br />

species to adapt to the contrasting soil conditions. The poster presents our experimental<br />

design and findings from preliminary collections made during fall <strong>of</strong><br />

2011 and spring <strong>of</strong> 2012.<br />

Jumpponen, Ari. Division <strong>of</strong> Biology, Kansas State University, Manhattan,<br />

KS66506. Tallgrass prairie soil microbial communities are resilient to climate<br />

change<br />

Climate models for central United States predict increasing temperatures<br />

and greater variability in precipitation. Combined, these shifts in environmental<br />

conditions are likely to impact many ecosystem properties and services. The existing<br />

Rainfall Manipulation Plots (RaMPs) experiment permits addressing address<br />

soil microbial community responses to simultaneous manipulation <strong>of</strong> temperature<br />

and temporal variability in precipitation. The RaMPs experiment is<br />

located in the tallgrass prairie at Konza Prairie Biological Station and has been operational<br />

since 1998 providing therefore a potential to address soil community responses<br />

to environmental manipulations in the long term. To test whether community<br />

composition, richness, or diversity respond to environmental change,<br />

more than 100,000 bacterial and fungal amplicons across 48 experimental units<br />

were 454-sequenced. The acquired data suggest that the soil communities are<br />

compositionally resilient to predicted environmental changes. This is the case<br />

both for the communities overall as inferred from ordination analyses as well as<br />

analyses <strong>of</strong> variance for each <strong>of</strong> the most common Operational Taxonomic Units<br />

(OTUs). However, while seemingly compositionally resilient, other data sources<br />

as well as on-going functional assays strongly indicate functional shifts in these<br />

communities.<br />

Justo, Alfredo. Clark University. Biology Department. 950 Main St, 01610,<br />

Worcester (MA). Lepiotaceous fungi (Agaricaceae, Basidiomycota) in Northeastern<br />

North <strong>America</strong><br />

Collections <strong>of</strong> Lepiotaceous fungi (Lepiota and allied genera in the Agaricaceae)<br />

made in the Northeastern U.S.A between 2009-2011 were studied using<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 25


morphological and molecular (nrITS) data. Two taxa are currently under revision<br />

and probably represent undescribed species: (i) Lepiota novae-angliae, nom.<br />

prov., recorded from Massachusetts and New Hampshire closely resembles L.<br />

clypeolaria in morphology but it’s ITS sequences do not match those <strong>of</strong> L. clypeolaria<br />

or any <strong>of</strong> the species <strong>of</strong> sect. Lepiota currently available in GenBank; (ii)<br />

Leucoagaricus allochthonus, nom. prov., was recorded in an indoors (greenhouse-like)<br />

garden in central Massachusetts and is probably a tropical alien<br />

species. Microscopically, it is characterized by the presence <strong>of</strong> pleurocystidia, a<br />

feature only shared by 3-4 other species in the very diverse Leucoagaricus/Leucocoprinus<br />

lineage (> 100 described species). Molecular data confirmed the presence<br />

<strong>of</strong> widespread cosmopolitan species in our area (e.g. Lepiota cristata, Leucoagaricus<br />

leucothites) but also revealed molecular differences between isolates<br />

<strong>of</strong> northeastern lepiotaceous fungi (e.g. Leucoagaricus rubrotinctus, Lepiota felina)<br />

and GenBank sequences deposited under the same name but with a different<br />

geographic origin.<br />

Kaffenberger, Justin T and Jonathan S Schilling. Department <strong>of</strong> Bioproducts<br />

and Biosystems Engineering, University <strong>of</strong> Minnesota, 2004 Folwell Avenue, St.<br />

Paul, MN 55108. Insights into the potential diversity <strong>of</strong> the brown rot decay<br />

mechanism provided by substrate compositional analysis<br />

Wood-degrading fungi are typically designated by the type <strong>of</strong> decay that<br />

they produce: the two main types being brown rot and white rot. While much effort<br />

has been put into understanding the biochemical mechanism by which brownrot<br />

occurs, the process remains poorly understood. Evidence suggests that brownrot<br />

fungi incorporate a Fenton reaction-based oxidation to rapidly de-polymerize<br />

substrate components and several mechanisms incorporating Fenton chemistry<br />

have been proposed. These proposed mechanisms have been based largely on<br />

studies performed on only a few brown-rot species. The polyphyletic nature <strong>of</strong><br />

brown-rot, however, suggests that there may be several routes to brown rot decay.<br />

While these fungi all ultimately produce brown decay residues, details regarding<br />

the chemical composition <strong>of</strong> these residues over the course <strong>of</strong> degradation have<br />

not been explored. Details regarding this chemistry may elucidate similarities and<br />

differences in the nature <strong>of</strong> underlying decay mechanisms. To this end, one representative<br />

from each <strong>of</strong> the known seven clades <strong>of</strong> brown rot fungi (Gloeophyllum<br />

trabeum, Fomitopsis pinicola, Ossicaulis lignatilis, Serpula lacrymans, Fistulina<br />

hepatica, Wolfiporia cocos, and Dacryopinax sp.) were used to decay<br />

representative hardwood (Populus tremuloides), s<strong>of</strong>twood (Pinus radiata), and<br />

grass (Zea mays) species. Compositional analyses <strong>of</strong> the residues were performed<br />

at six time points over the course <strong>of</strong> 4 months <strong>of</strong> exposure. Details <strong>of</strong> these compositions<br />

and their implications are discussed.<br />

Kauserud, Håvard 1 , Einar Heegaard 2 , Ulf Büntgen 3 , Simon Egli 3 , Rune<br />

Halvorsen 4 , and Lynne Boddy 5 . 1 Microbial Evolution Research Group (MERG),<br />

Department <strong>of</strong> Biology, University <strong>of</strong> Oslo, PO Box 1066 Blindern, NO-0316<br />

Oslo, Norway, 2 Norwegian Forest and Landscape Institute, Fanaflaten 4, N-5244<br />

Fana, Norway, 3 Swiss Federal Research Institute for Forest Snow and Landscape<br />

(WSL), Birmensdorf, Switzerland, 4 Department <strong>of</strong> Research and Collections,<br />

Natural History Museum, University <strong>of</strong> Oslo, PO Box 1172 Blindern, NO-0318<br />

Oslo, Norway, 5 Cardiff School <strong>of</strong> Biosciences, Biomedical Building, Museum<br />

Avenue, Cardiff CF10 3AX, UK. Temporal patterns <strong>of</strong> fungal fruiting reveal<br />

ongoing climate change effects<br />

Effects <strong>of</strong> climate change on fungi may manifest in various ways. Time<br />

<strong>of</strong> fruiting may change, yields <strong>of</strong> fruit bodies can be altered and distributional<br />

ranges <strong>of</strong> species may shift. We are studying these three aspects using statistical<br />

analyses <strong>of</strong> time series data and digitized field and herbarium records. We show<br />

that the annual fruiting period across Europe has changed during the last 50 years.<br />

In Austria, Norway, Switzerland and the UK, the average annual fruiting time <strong>of</strong><br />

autumn fruiters has been delayed. Moreover, the start and end <strong>of</strong> the annual fruiting<br />

period, using the 5 and 95 percentiles as proxies, have changed towards earlier<br />

and later, respectively. Changes in fruiting time have been especially dramatic<br />

in the UK, which deviates from the other regions by a more oceanic climate. Time<br />

series data from a Swiss forest plot indicates that the annual yields <strong>of</strong> ectomycorrhizal<br />

fruit bodies has increased dramatically during the last 40 years. We hypothesize<br />

that changes in the regional climate, including higher temperatures and<br />

more precipitation, can be linked to the increased fruit body yields. In another ongoing<br />

study we are exploring the extent to which fungi in the UK have shifted distributional<br />

ranges during the last 50 years and whether this can be linked to climate<br />

change.<br />

Kepler, Ryan M 1 , Jonathan Shao 2 , and Stephen A Rehner 1 . 1 Systematic Mycology<br />

2<br />

and Microbiology Laboratory, USDA, ARS, Beltsville, MD, 20705,<br />

Molecular Plant Pathology Laboratory, USDA, ARS, Beltsville, MD, 20705.<br />

Genome-enabled marker development for evolutionary studies <strong>of</strong> the insect<br />

pathogen Metarhizium<br />

Genome sequence resources for fungi enable development <strong>of</strong> custom genetic<br />

markers scaled to specific evolutionary inferences that improve greatly on<br />

26 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

information derived from the more limited set <strong>of</strong> generic legacy genetic markers.<br />

Using recently available genome sequences for M. anisopliae and M. acridum, we<br />

have developed a suite <strong>of</strong> novel genetic markers for investigating cryptic speciation,<br />

phylogeography and population genetics within Metarhizium s.l. To exploit<br />

the inherent variability <strong>of</strong> nuclear intergenic (nucIGS) regions we have created a<br />

simple graphical approach to visually scan for variable nucIGS regions flanked by<br />

conserved gene sequences that can serve as anchors for robust primer development.<br />

Using pairwise alignments <strong>of</strong> selected genomic scaffolds from the two<br />

species created in the Mummer 3.0 genome alignment program, a simple script<br />

was developed that tabulates the number <strong>of</strong> SNPs via a scalable sliding window<br />

function, facilitating visualization <strong>of</strong> trends in nucleotide substitutions and similarity<br />

between syntenic genomic regions among pairs <strong>of</strong> sequences. We describe<br />

the development <strong>of</strong> novel markers developed using this approach and demonstrate<br />

their utility to infer cryptic phylogenetic partitions previously missed or poorly resolved<br />

by pre-existing phylogenetic markers. We demonstrate the superior performance<br />

<strong>of</strong> these new versus preexisting markers by comparison <strong>of</strong> their phylogenetic<br />

informativeness using PhyDesign and Genealogical Sorting Indices. In<br />

addition we: 1) summarize the numbers, types and genomic distribution <strong>of</strong> microsatellite<br />

markers and compare these to those developed by enrichment methods,<br />

and 2) discuss the mating locus and development <strong>of</strong> improved mating type<br />

assignment PCR assays. Through selected examples we illustrate the improvements<br />

in evolutionary inference achieved with these newly developed genetic<br />

markers.<br />

Kerekes, Jennifer F 1 , Michael Kaspari 2 , Bradley S Stevenson 3 , and Thomas D<br />

Bruns 1 . 1 111 Koshland Hall, Dept. <strong>of</strong> Plant and Microbial Biology, UC Berkeley,<br />

Berkeley CA 94720, USA, 2 Graduate Program in EEB, Dept. <strong>of</strong> Zoology,<br />

University <strong>of</strong> Oklahoma, Norman, OK 73019, USA, 3 Graduate Program in EEB,<br />

Dept. <strong>of</strong> Botany and Microbiology, University <strong>of</strong> Oklahoma, Norman, OK 73019,<br />

USA. Nutrient enrichment increased species richness <strong>of</strong> leaf litter fungal<br />

communities in a diverse lowland tropical forest<br />

We explored leaf litter fungal diversity in a diverse lowland tropical forest<br />

in which a replicated factorial N, P, K, micronutrient fertilization experiment<br />

<strong>of</strong> 40 x 40 m plots had been on going for nine years. Fertilized plots were compared<br />

with control plots that did not receive fertilization to evaluate possible nutrient<br />

effects on leaf litter fungal communities. We extracted DNA from leaf litter<br />

samples collected on the Gigante peninsula, Barro Colorado Nature Monument<br />

(BCNM), Republic <strong>of</strong> Panama and used fungal-specific amplification and a 454<br />

pyrosequencing approach to sequence two loci, the nuclear ribosomal internal<br />

transcribed spacer (ITS) region and the nuclear ribosomal large subunit (LSU) D1<br />

region. Overall, both the ITS and LSU gene regions had similar results. The use<br />

<strong>of</strong> two separate gene regions for the analysis provided a complimentary look at<br />

the data. Ascomycota (including approximately 29 orders, 44 families, and 88<br />

genera) are the dominant phylum among the leaf litter fungi, followed by Basidiomycota<br />

(including approximately 31 orders, 38 families and 52 genera). The<br />

long-term addition <strong>of</strong> nutrients increased species richness relative to the control<br />

plots, and had an effect on the taxonomic composition <strong>of</strong> the leaf litter fungal<br />

communities at lower taxonomic levels (i.e. family, genus and species), but not at<br />

higher taxonomic levels (i.e. phylum, class and order). Nitrogen has the largest effect,<br />

followed by phosphorus and potassium.<br />

Kerrigan, Julia and Virginia Waldrop. School <strong>of</strong> Agricultural, Forest, and Environmental<br />

Sciences, 114 Long Hall, Clemson University, Clemson, SC 29<strong>63</strong>4-<br />

0310. Bi<strong>of</strong>ilm formation by Aspergillus niger, Aureobasidium pullulans, and<br />

Cladosporium cladosporioides<br />

A bi<strong>of</strong>ilm is an aggregate <strong>of</strong> microorganisms adhered together, to each<br />

other and to a substrate, surrounded by an extracellular matrix. Bi<strong>of</strong>ilms are ubiquitous,<br />

occurring in natural and artificial water systems, and are <strong>of</strong> great concern<br />

because <strong>of</strong> their persistence in industrial, medical, and household environments.<br />

Although bacterial bi<strong>of</strong>ilm research is well established, fungal bi<strong>of</strong>ilm research is<br />

comparatively new and the vast majority <strong>of</strong> the research has been conducted on<br />

yeasts. This research was undertaken to develop a repeatable system for studying<br />

bi<strong>of</strong>ilm formation by different types <strong>of</strong> filamentous fungi and to document their<br />

phenotypic changes throughout bi<strong>of</strong>ilm development. Aspergillus niger, Aureobasidium<br />

pullulans, and Cladosporium cladosporioides were studied because<br />

they are cosmopolitan, commonly found in bi<strong>of</strong>ilms, and have characteristic culture<br />

morphology allowing for easy detection <strong>of</strong> contaminants. We established a<br />

method for engineering bi<strong>of</strong>ilms in a controlled reactor that models those growing<br />

under low sheer. This method can be modified for different types <strong>of</strong> filamentous<br />

fungi and other microbes. The first stage <strong>of</strong> filamentous fungal bi<strong>of</strong>ilm formation<br />

involves spore attachment and germination. Hyphal tip growth produces<br />

an extracellular polysaccharide that helps the fungus grow and allows for attachment<br />

within the liquid environment. Hyphae form an interconnected network that<br />

attaches to itself and other surfaces. As hyphal proliferation continues, spores are<br />

produced, germinate, and give rise to additional bi<strong>of</strong>ilm mass. A mature bi<strong>of</strong>ilm<br />

Continued on following page


is composed <strong>of</strong> a confluent mucilaginous hyphal matrix covered with conidia;<br />

however, the morphological changes during bi<strong>of</strong>ilm formation vary substantially<br />

among the species. The establishment <strong>of</strong> a repeatable filamentous fungus bi<strong>of</strong>ilm<br />

formation system and understanding the morphological steps involved in bi<strong>of</strong>ilm<br />

formation allow for future studies on how to prevent and remove them.<br />

Kim, Seong Hwan 1 , Hyuk Woo Kwon 1 , Min Woo Hyun 1 , Gi-Ho Sung 2 , and<br />

Hyung-Kyoon Choi 3 . 1 Department <strong>of</strong> Microbiology, Dankook University, Cheonan,<br />

Chungnam 330-714, Korea, 2 2Mushroom Research Division, National Institute<br />

<strong>of</strong> Horticultural and Herbal Science, Rural Development Administration,<br />

Suwon 441-707, Korea, 3 College <strong>of</strong> Pharmacy, Chung-Ang University, Seoul<br />

156-756, Korea. Microscopic analysis <strong>of</strong> Cordyceps bassiana (anamorph<br />

stage: Beauveria bassiana) stromata during artificial cultivation for commercial<br />

Use<br />

Cordyceps bassianais is an insect-borne medicinal fungus that has recently<br />

been cultivated artificially. To determine the proper time for harvesting<br />

fruit bodies <strong>of</strong> C. bassiana with commercial value, stereo and scanning electron<br />

microscopic images <strong>of</strong> perithecium formation on stomata were analyzed during<br />

13 weeks <strong>of</strong> artificial cultivation. Perithecia formation initiated on the stromata<br />

after 5 weeks in culture. After 6 weeks they were nearly mature on the stromata<br />

but ascospores had not been released. Fully matured perithecia producing ascospores<br />

were found on the stromata after 7 weeks. The best looking morphological<br />

features <strong>of</strong> fruit bodies with good commercial value were found after 6 to 7<br />

weeks culture. The highest level <strong>of</strong> total phenolics content was obtained from fruit<br />

bodies after 6 weeks culture. Our work suggests the formation and degree <strong>of</strong><br />

perithecia maturity could be determined according to fruit body development and<br />

applied as a criterion for harvest time determination in the production <strong>of</strong> secondary<br />

metabolites from artificially grown C. bassiana stromata.<br />

Kluting, Kerri L 1 , Sarah E Bergemann 1 , and Timothy J Baroni 2 . 1 Middle Tennessee<br />

State University, Biology Department, PO Box 60, Murfreesboro TN<br />

37132, 2 State University <strong>of</strong> New York, College at Cortland, Department <strong>of</strong> Biological<br />

Sciences, PO Box 2000, Cortland NY 13045. Resolving the generic<br />

boundaries within the Rhodocybe/Clitopilus clade (Agaricomycetes, Basidiomycota)<br />

The Entolomataceae (Agaricomycetes, Basidiomycota) is a highly diverse<br />

assemblage <strong>of</strong> euagarics with well over 1000 taxa worldwide. These macr<strong>of</strong>ungi<br />

are recognized as a cohesive unit by their pink or flesh brown basidiospores<br />

that are angular in polar view, have some form <strong>of</strong> a ridged or pustulate ornamentation<br />

and the walls <strong>of</strong> the spores are evenly cyanophilic. Recent published phylogenetic<br />

analyses support the separation <strong>of</strong> these taxa into two clades, typically<br />

referred to as the ‘Entoloma’ and ‘Rhodocybe/Clitopilus’ clades. The<br />

Rhodocybe/Clitopilus clade is the focus <strong>of</strong> our investigation, as a recent publication<br />

suggests that the genus Rhodocybe (pustulate spores) should be subsumed<br />

under the genus Clitopilus (longitudinally ridged spores) in order to form a monophyletic<br />

clade. Our goal is to conduct a systematic revision <strong>of</strong> the Rhodocybe/Clitopilus<br />

clade, focusing on delimitation <strong>of</strong> generic boundaries using a phylogenetic<br />

framework. Initial efforts have focused on obtaining sequences <strong>of</strong><br />

mitochondrial and low-copy nuclear loci [mitochondrial small subunit ribosomal<br />

RNA (mtSSU) and ATPase subunit 6 (ATP6), the nuclear RNA polymerase second<br />

largest subunit (RPB2) and the translation elongation factor alpha gene<br />

(TEF)] to define the generic boundaries. To date, we have sequenced and performed<br />

a phylogenetic analysis utilizing 86 collections <strong>of</strong> Clitopilus and<br />

Rhodocybe. The results <strong>of</strong> our combined analysis <strong>of</strong> the four genes provide significant<br />

support for the delineation <strong>of</strong> the following clades within Rhodocybe: i) a<br />

clade containing Clitopilopsis hirneola (formerly as Rhodocybe hirneola); ii) a<br />

fallax/mundula/popinalis clade nested between Clitopilus and Clitopilopsis; iii) a<br />

Rhodocybe s.s. clade with several Rhodocybe taxa, including the type for the<br />

genus, R. caelata and; iv) a clade containing R. nitellina. A discussion <strong>of</strong> possible<br />

scenarios to explain evolution in the Rhodocybe/Clitopilus clade will be presented.<br />

Koch, Rachel A 1 , Terry W Henkel 2 , and M Catherine Aime 1 . 1 Department <strong>of</strong><br />

Plant Pathology and Crop Physiology, Louisiana State University Agricultural<br />

Center, Baton Rouge, Louisiana 70803 USA, 2 Department <strong>of</strong> Biological Sciences,<br />

Humboldt State University, Arcata, California 95521 USA. Resolving the<br />

phylogenetic relationship between Armillaria and the newly discovered<br />

genus Guyanagaster<br />

Recent mycological work in the remote Pakaraima Mountains <strong>of</strong> Guyana<br />

uncovered Guyanagaster necrorhiza, a new gasteromycete genus and species in<br />

the Physalacriaceae (Agaricales). The original study showed that Guyanagaster is<br />

most closely related to the agaricoid genus Armillaria and has a similar root parasitism/wood<br />

decay habit. While Armillaria species exhibit typical lamellate<br />

mushroom morphologies Guyanagaster basidiomata are entirely sequestrate with<br />

a thick, hard peridium, tough, decay-resistant gleba with highly modified hymenium<br />

and heavily ornamented basidiospores, and a short stipe attached directly to<br />

root substrata. Molecular phylogenetic analyses using different gene regions and<br />

varying numbers <strong>of</strong> taxa were, however, not in complete agreement as to the exact<br />

relationship between Guyanagaster and Armillaria; some analyses suggested that<br />

Guyanagaster is sister to Armillaria, whereas others indicated that Guyanagaster<br />

is nested within Armillaria. To better understand the phylogenetic relationships <strong>of</strong><br />

Guyanagaster and Armillaria within the Physalacriaceae as well as implications<br />

for evolution <strong>of</strong> basidioma forms, we built an expanded three-gene supermatrix<br />

(ITS, LSU and EF1a) from an increased number <strong>of</strong> Armillaria species, Guyanagaster,<br />

and other exemplar Physalacriaceae taxa. A maximum likelihood analysis<br />

<strong>of</strong> these data suggested that Guyanagaster may be derived from within Armillaria,<br />

a result consistent with demonstrated origins <strong>of</strong> sequestrate taxa within other<br />

lamellate lineages. In light <strong>of</strong> these results we feel that retaining generic status for<br />

Guyanagaster is warranted due to its extreme morphological divergence from<br />

Armillaria; such an approach will, however, render Armillaria sensu lato paraphyletic.<br />

We also present a new species <strong>of</strong> Armillaria that occurs sympatrically<br />

with Guyanagaster and is the first record <strong>of</strong> an Armillaria species in Guyana.<br />

Kozanitas, Melina, Todd W Osmundson, and Matteo Garbelotto. Forest Mycology<br />

and Pathology Laboratory, Department <strong>of</strong> Environmental Science, Policy and<br />

Management, University <strong>of</strong> California, Berkeley, CA 94720, USA. Temporal<br />

fluctuations and the role <strong>of</strong> disturbance in disease progression <strong>of</strong> the Sudden<br />

Oak Death epidemic<br />

With its high host mortality and ability to cause landscape-scale alterations<br />

in forest cover and composition, Sudden Oak Death (etiological agent<br />

Phytophthora ramorum, Stramenopila, Oomycota) mirrors past forest disease<br />

epidemics such as Chestnut Blight and Dutch Elm Disease. In contrast with these<br />

past epidemics, however, the appearance <strong>of</strong> Sudden Oak Death converges with a<br />

time <strong>of</strong> significant advancement in the development <strong>of</strong> molecular genetic tools<br />

that allow the movement <strong>of</strong> individuals (or individual genotypes) to be tracked<br />

and the role <strong>of</strong> evolutionary processes in disease progression to be assessed. Such<br />

methods have been instrumental in reconstructing the likely origins and geographic<br />

pathways <strong>of</strong> spread <strong>of</strong> the pathogen; however, little is yet known about<br />

the local-scale processes contributing to disease maintenance and progression.<br />

Since 2008, we have surveyed a network <strong>of</strong> sampling plots in the San Francisco<br />

Bay Area multiple times per year in order to examine seasonal patterns in the incidence<br />

and population genetic structure <strong>of</strong> infection foci. A period <strong>of</strong> severe<br />

drought in 2007-2009 followed by a period <strong>of</strong> normal to above-normal precipitation<br />

in 2010-2011 provide a significant opportunity to examine the effects <strong>of</strong> abiotic<br />

disturbance on disease progression. Here, we present recent and ongoing research<br />

focused on assessing seasonal patterns in genotypic diversity <strong>of</strong> viable<br />

infections that can act as reservoirs <strong>of</strong> new infectious propagules, assessing the effect<br />

<strong>of</strong> drought as a potential agent <strong>of</strong> selection on pathogen genotypes, examining<br />

the relationship between infections on dead-end (Oak) and amplifying (California<br />

Bay Laurel) hosts, and assessing the possible role <strong>of</strong> competitive<br />

interactions with sympatric Phytophthora species (P. pseurosyringae, P.<br />

nemorosa) using culture-based surveys, culture-independent (qPCR) assays, and<br />

analyses <strong>of</strong> population genetic structure (based on variable microsatellite loci) to<br />

infer the underlying processes <strong>of</strong> pathogen demographic expansion, contraction<br />

and spread.<br />

Krüger, Dirk 1 , Björn Hoppe 1 , Witoon Purahong 1 , Danuta Kapturska 1 , Kristin<br />

Baber 2 , Marek Pecyna 3 , Tobias Arnstadt 3 , Renate Rudl<strong>of</strong>f 1 , Harald Kellner 3 ,<br />

Peter Otto 2 , Tiemo Kahl 4 , François Buscot 1 , Jürgen Bauhus 4 , and Martin<br />

H<strong>of</strong>richter 3 . 1 UFZ - Helmholtz Centre for Environmental Research, Department<br />

<strong>of</strong> Soil Ecology, Th.-Lieser-Str. 4, D-06120 Halle/Saale, Germany, 2 University<br />

<strong>of</strong> Leipzig, Institute <strong>of</strong> Biology I, Herbarium Universitatis Lipsiensis (LZ), Johannisallee<br />

21-23, D-04103 Leipzig, Germany, 3 International Graduate School<br />

Zittau – IHI, Markt 23, D-027<strong>63</strong> Zittau, Germany, 4 University <strong>of</strong> Freiburg, Institute<br />

<strong>of</strong> Silviculture, Tennenbacherstr. 4, D-79085 Freiburg i. Brsg., Germany.<br />

Overview <strong>of</strong> research activities <strong>of</strong> the FunWood and FuPerS project consortia<br />

<strong>of</strong> the German Biodiversity Exploratories: fungal and bacterial diversity<br />

and wood and litter decay<br />

In forests ecosystems, coarse woody debris (CWD, a.k.a. dead wood) and<br />

litter are dominant habitats rich in lignin and cellulose. Their degradation, mainly<br />

by Fungi, is an important part <strong>of</strong> nutrient cycling. Intensive use <strong>of</strong> forests for timber<br />

production increases loss <strong>of</strong> CWD, thus may affect the diversity <strong>of</strong> organisms<br />

sharing this habitat. The FunWood project <strong>of</strong> the DFG (German Science Foundation)<br />

Biodiversity Exploratories investigates the postulated effect <strong>of</strong> forest management<br />

intensity on the diversity <strong>of</strong> wood-decaying fungi and dead wood decomposition.<br />

In the first project phase the fungal diversity was approached both<br />

by molecular biological methods (fingerprinting and sequencing <strong>of</strong> clone libraries)<br />

and surveying fruit bodies. In addition, preliminary investigations on bacterial<br />

diversity, including nitrogen fixation genes (nifH), were performed. Potentially<br />

new bacteria were identified by nifH. The follow-up, second phase <strong>of</strong> this<br />

project (started 2011) scales up the study on wood decay fungi to encompass 150<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 27


plots in all three exploratory locations (Schwäbische Alb, Hainich-Dün,<br />

Schorfheide-Chorin). Thirteen species <strong>of</strong> trees whose logs were placed there as<br />

the BeLongDead experiment are now included. A two-year experiment <strong>of</strong> adding<br />

nitrogen to dead wood was also initiated. We discuss some results <strong>of</strong> phase 1 and<br />

some planned activities in phase 2. In addition, some current aspects <strong>of</strong> a related<br />

project, FuPerS, studying leaf litter degradation, are discussed.<br />

LeBlanc, Nicholas, Linda Kinkel, and H Corby Kistler. Department <strong>of</strong> Plant<br />

Pathology, University <strong>of</strong> Minnesota, 495 Borlaug Hall 1991 Upper Buford Circle,<br />

St. Paul, MN 55108. Soil fungal lineages respond to host plant identity and<br />

community diversity in a model grassland system<br />

Grassland soils are known to harbor diverse fungal communities but the<br />

effects <strong>of</strong> host plant identity and the diversity <strong>of</strong> the surrounding plant community<br />

on these fungi is unknown. We have focused on rhizosphere fungal communities<br />

in a model grassland system using targeted metagenomics and well-annotated<br />

reference material for taxonomic assignment. A total <strong>of</strong> twenty-four samples<br />

were analyzed from hosts grown in monoculture or polyculture. Hosts included<br />

Big Bluestem, Little Bluestem, Round-Headed Bush Clover, and Sundial Lupine.<br />

We hypothesized that host family (grasses and legumes) and plant community diversity<br />

(monoculture and polyculture) would influence fungal community compositions<br />

and diversity. Amplicon libraries, derived from rhizosphere DNA, were<br />

sequenced using 454 technology. Sequences were aligned to a custom database <strong>of</strong><br />

4030 unique accessions published in phylogenetic studies, derived from genome<br />

databases, and obtained from the AFTOL consortium. Initial NCBI taxon IDs<br />

were used to assign reads to higher taxonomies using Perl scripts. Approximately<br />

800 fungal genera were detected from the prairie soils. More diverse fungal<br />

communities were associated with hosts grown in polyculture. Many <strong>of</strong> the assigned<br />

taxonomies represent slow-growing, poorly- studied fungi, such as<br />

Chaetothyriales and Verrucariales, which were enriched in the grass rhizosphere.<br />

Other taxa, like Helotiales, were preferentially enriched in more diverse plant<br />

communities but also significantly correlated with soil potassium. This work<br />

highlights the first steps towards linking plant host and community diversity with<br />

the abundance <strong>of</strong> root-associated fungi that likely play important, but poorly understood<br />

roles in plant growth and development.<br />

Lehr, Nina A 1 , Zheng Wang 1 , Francesc Lopez-Giraldez 1 , Ning Li 1 , Frances<br />

Trail 2 , and Jeffrey Townsend 1 . 1 2<br />

Yale University, New Haven, CT, USA,<br />

Michigan State University, East Lansing, MI, USA. The importance <strong>of</strong><br />

carotenoids for sexual development <strong>of</strong> fungi<br />

Carotenoids are terpenes which belong to the most widespread group <strong>of</strong><br />

compounds in nature and which are produced by plants, bacteria and fungi. Terpenes<br />

fulfill diverse functions such as protection against reactive oxygen species<br />

(ROS) and are beneficial for the nutrition <strong>of</strong> humans and animals. The orange pigmentation<br />

<strong>of</strong> the fungus Neurospora crassa derives from the accumulation <strong>of</strong> the<br />

xanthophyll neurosporaxanthine and other precursor carotenoids in response to<br />

asexual spore development upon exposure to light. Carotenoid biosynthesis in N.<br />

crassa has been studied intensively and it is known that blue light controls induction<br />

<strong>of</strong> carotenoid production in the mycelium, formation <strong>of</strong> protoperithecia as<br />

well as phototropism <strong>of</strong> perithecial beaks and perithecial polarity. Carotenoid<br />

biosynthesis involves five different enzymes, amongst which are three “albino”<br />

genes al-1, al-2 and al-3, encoding enzymes essential for carotenogenesis but not<br />

for growth. In its life cycle, Neurospora undergoes either an asexual or a sexual<br />

phase depending on environmental conditions. The asexual phase results in hyphal<br />

growth leading to the formation <strong>of</strong> macro- and microconidia while in the sexual<br />

phase complex, three-dimensional fruiting bodies, perithecia, are formed. So<br />

far there is no information on the linkage between carotenoid biosynthesis and the<br />

sexual life cycle. We have performed Illumina next generation sequencing <strong>of</strong><br />

three Neurospora species over the time course <strong>of</strong> sexual development, followed<br />

by a comparative gene expression analysis with a pipeline we have generated to<br />

measure the levels <strong>of</strong> gene expression. We have quantified the amount <strong>of</strong> total<br />

carotenoids present in the mycelium at each individual time point. A screen <strong>of</strong><br />

knockout mutants has revealed a linkage <strong>of</strong> carotenoid biosynthesis to the formation<br />

<strong>of</strong> perithecia.<br />

Lehr, Nina 1 , Zheng Wang 1 , Usha Sikhakolli 2 , Francesc Lopez-Giraldez 1 , Ning<br />

Li 1 , Frances Trail 2 , and Jeffrey P Townsend 1 . 1 Department <strong>of</strong> Ecology and<br />

Evolutionary Biology and Program in Computational Biology and Bioinformatics,<br />

Yale University, New Haven, Connecticut 06520, USA., 2 Department <strong>of</strong><br />

Plant Biology, Michigan State University, East Lansing, MI 48824. Comparative<br />

transcriptomics reveals new Neurospora crassa genes important to<br />

perithecial development<br />

In recent years, a plethora <strong>of</strong> genomic sequences have been released for<br />

fungal species, accompanied by functional predictions for genes based on protein<br />

sequence comparisons. However, identification <strong>of</strong> genes involved in particular<br />

processes has been extremely slow, and new methodologies for identifying genes<br />

involved in a particular process have not kept pace with the exponential increase<br />

28 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

in genome sequence availability. We have performed transcriptional pr<strong>of</strong>iling <strong>of</strong><br />

five species <strong>of</strong> Neurospora and Fusarium during six stages <strong>of</strong> perithecium development.<br />

Because we maintained a strictly common medium across experiments,<br />

our transcriptomic data revealed solely evolved differences in the transcriptional<br />

basis <strong>of</strong> morphological changes. We estimated ancestral gene expression pr<strong>of</strong>iles<br />

and transcriptional shifts across this developmental process, facilitating identification<br />

<strong>of</strong> genes whose transcription had substantially and significantly shifted during<br />

the evolutionary process. We examined one hundred genes whose expression<br />

greatly increased in Neurospora crassa perithecial development compared to<br />

Neurospora tetrasperma, compared to Neurospora discreta, or compared to<br />

Fusarium spp. Phenotypes <strong>of</strong> knockouts <strong>of</strong> these genes included substantial<br />

changes in the timing and environmental sensitivity <strong>of</strong> perithecial development.<br />

These genes were not previously identified as candidates for function in perithecium<br />

development, illustrating the utility <strong>of</strong> this method for identification <strong>of</strong> genes<br />

associated with specific functional processes.<br />

Lendemer, James C 1 and Brendan P Hodkinson 2 . 1 Institute <strong>of</strong> Systematic<br />

Botany, The New York Botanical Garden, Bronx, NY 10458-5126, U.S.A., 2 International<br />

Plant Science Center, The New York Botanical Garden, Bronx, NY<br />

10458-5126, USA. Sterile asexually reproducing crustose lichens: improving<br />

classification, conservation, and communication in the 21st century<br />

Although many lichens reproduce through the dispersal <strong>of</strong> sexual diaspores,<br />

asexual reproduction is not uncommon and occurs in nearly all <strong>of</strong> the diverse<br />

lineages that comprise lichen-forming fungi. This type <strong>of</strong> reproduction can<br />

occur through either purely fungal diaspores or lichenized diaspores. The latter<br />

comprise a morphologically diverse array <strong>of</strong> structures that have evolved to facilitate<br />

the co-dispersal <strong>of</strong> the components <strong>of</strong> the lichen microbiome. The majority<br />

<strong>of</strong> crustose lichens that reproduce exclusively via lichenized diaspores are poorly<br />

understood, under collected, or undescribed and they represent a large component<br />

<strong>of</strong> lichen biodiversity that is currently overlooked and underestimated. We present<br />

a conceptual summary <strong>of</strong> the results <strong>of</strong> five years <strong>of</strong> research on this group in<br />

the context <strong>of</strong> a revision <strong>of</strong> the sterile crustose genus Lepraria s.l. In addition to<br />

extensive morphological studies <strong>of</strong> thallus ultrastructure, and integrated studies <strong>of</strong><br />

molecular and non-molecular datasets to resolve species boundaries, we show that<br />

Lepraria s.l. is polyphyletic at the ordinal level and contains disparate elements<br />

belonging to four families in three orders, one <strong>of</strong> which was previously unrecognized.<br />

Drawing upon this body <strong>of</strong> work we present a standard protocol to improve<br />

our understanding <strong>of</strong> sterile crustose lichens, communicate this understanding to<br />

a diverse array <strong>of</strong> interdisciplinary researchers, and translate this understanding<br />

into meaningful applications to promote real-world conservation.<br />

Letcher, Peter M, Satoshi Sekimoto, and Martha J Powell. Department <strong>of</strong> Biological<br />

Sciences, The University <strong>of</strong> Alabama, Tuscaloosa, AL 35487. Characterizing<br />

the paracrystalline inclusion in zoospores <strong>of</strong> members <strong>of</strong> the<br />

Chytridiales (Chytridiomycota)<br />

Orders within Chytridiomycota (=chytrids) have now been circumscribed<br />

as monophyletic lineages, each with zoospores exhibiting unique suites <strong>of</strong> character<br />

states. Molecular phylogenetic analyses allow us to hypothesize zoospore<br />

character evolution through tracking modifications, losses, or appearances <strong>of</strong><br />

characters along these lineages. The presence <strong>of</strong> a paracrystalline inclusion is one<br />

zoosporic character within Chytridiomycota that has been reported only among<br />

members <strong>of</strong> the Chytridiales. We are investigating the structure and function <strong>of</strong><br />

the paracrystalline inclusion in zoospores <strong>of</strong> the Chytridiales. Although the<br />

Chytridiales includes a variety <strong>of</strong> zoospore sub-types defining the families<br />

Chytridiaceae and Chytriomycetaceae, zoospores <strong>of</strong> all members studied with<br />

electron microscopy contain a paracrystalline inclusion. Consequently, the presence<br />

<strong>of</strong> a paracrystalline inclusion unites the Chytridiales and serves as a diagnostic<br />

character for this order within Chytridiomycota. Because paracrystalline inclusions<br />

found in organisms outside the Chytridiomycota are typically associated<br />

with the accumulation <strong>of</strong> proteins, our hypothesis is that the paracrystalline inclusion<br />

in zoospores <strong>of</strong> Chytridiales is composed <strong>of</strong> protein. We have found that<br />

number, size, and complexity <strong>of</strong> paracrystalline inclusions vary with species. The<br />

long-term aim <strong>of</strong> this study is to isolate the paracrystalline inclusion and characterize<br />

it biochemically.<br />

Levesque, C Andre, Gregg Robideau, Chen Wen, and Lewis Christopher. Agriculture<br />

& Agri-Food Canada, Ottawa, Canada. Oomycete barcoding and applications<br />

The ITS region is the de facto DNA barcode in oomycetes as in true<br />

Fungi. However, the cytochrome oxidase 1 (COI) region is the NCBI accepted<br />

barcode region and it must be ruled out to pave the way for an alternative marker<br />

such as ITS. COI does not work for the true Fungi, primarily because <strong>of</strong> numerous<br />

introns, but in oomycetes COI works at least as well as ITS. The prevalence<br />

<strong>of</strong> ITS in mycology has led to the proposal that it be included as barcode in addi-<br />

Continued on following page


tion to COI for oomycetes. Recently, the oomycete COI database was expanded<br />

to have comparable coverage <strong>of</strong> oomycetes species to the existing ITS data. COI<br />

has the advantage <strong>of</strong> being easy to align across all oomycetes because there are no<br />

introns and indels, whereas it is impossible to align ITS between genera. However,<br />

ITS was found to be a more valuable target in a DNA array system developed<br />

for most Phytophthora species. Specificity <strong>of</strong> hybridization probes is increased<br />

when they fall within an insertion or span deletions in closely related species. On<br />

the other hand, changes at the synonymous third codon positions are not as effective<br />

as they <strong>of</strong>ten yield very few base pair differences between closely related<br />

species over short oligonucleotides. Next generation sequencing provides additional<br />

applications <strong>of</strong> barcode sequence databases. The most common universal<br />

primers for ITS do amplify oomycete DNA and generate oomycete sequences in<br />

the large pool <strong>of</strong> PCR products processed by pyrosequencing. However, when<br />

studying ecology <strong>of</strong> oomycetes, the use <strong>of</strong> oomycete specific primers provides a<br />

better pr<strong>of</strong>ile <strong>of</strong> species that are present at low propagule concentration.<br />

Li, Ning and Jeffrey P Townsend. Department <strong>of</strong> Ecology and Evolutionary Biology<br />

and Program in Computational Biology and Bioinformatics, Yale University,<br />

New Haven, Connecticut 06520, USA. Estimating selective pressure<br />

across the coding sequences <strong>of</strong> the model yeast Saccharomyces cerevisiae<br />

The impact <strong>of</strong> natural selection on the genome has been well studied in<br />

diverse model species. However, genome-wide scans <strong>of</strong> polymorphism and divergence<br />

in synonymous and replacement sites within genes have yielded surprisingly<br />

few genes whose tallies reject a null hypothesis, especially in organisms<br />

with very high effective population sizes such as the model yeast Saccharomyces<br />

cerevisiae. The surprising dearth may arise because targets <strong>of</strong> selection are <strong>of</strong>ten<br />

likely to be restricted to a single functional domain or a few local, interacting sites<br />

within a protein. Unfortunately, most data sets lack sample size under previous<br />

methodologies to effectively detect selection at these levels, leading to a tyranny<br />

<strong>of</strong> the null hypothesis. Additionally, local variation in mutation rate <strong>of</strong> synonymous<br />

sites presents a challenge for fine-scale detection <strong>of</strong> a history <strong>of</strong> natural selection.<br />

To address these challenges, we have developed an approach that performs<br />

model-averaged clustering <strong>of</strong> intragenic polymorphism and divergence.<br />

Applying Poisson Random Field theory to the model-averaged polymorphism<br />

and divergence over all potential clusters, we graphically pr<strong>of</strong>ile the model-averaged<br />

level <strong>of</strong> selection and its 95% confidence intervals for each site. Our genomic<br />

analysis <strong>of</strong> genes across the S. cerevisiae genome demonstrates that although<br />

much <strong>of</strong> the coding sequence is under purifying rather than adaptive<br />

natural selection, nearly 50% <strong>of</strong> coding sequence overall has been evolving under<br />

positive selection. Moreover, adaptively evolving sites are distributed across<br />

about 80% <strong>of</strong> the genes in the whole genome. The presence <strong>of</strong> extensive regions<br />

<strong>of</strong> purifying selection explains the finding that few genes are significant under traditional<br />

tests <strong>of</strong> polymorphism and divergence, even though many critical sites or<br />

regions within individual genes are playing important adaptive roles in an evolutionary<br />

history <strong>of</strong> adaptive natural selection.<br />

Lim, Young Woon 1 , Hyun Lee 1 , Changmu Kim 2 , and Jin Sung Lee 2 . 1 School<br />

<strong>of</strong> Biological Science, Seoul National University, Seoul, 151-747, Korea, 2 National<br />

Institute <strong>of</strong> Biological Resources (NIBR), Incheon, 404-708, Korea. Fungal<br />

diversity associated with oaks in Korea<br />

Eleven species <strong>of</strong> oak trees are common in Korean forests in association<br />

with evergreen pines. Although they have great economic value, oak trees have<br />

received little attention. With the goal to manage oak forests effectively, we investigated<br />

the diversity <strong>of</strong> decay-causing fungi and ectomycorrhizal (EM) fungi<br />

from the eight species <strong>of</strong> oak trees. Through examination <strong>of</strong> the specimens deposited<br />

in the Seoul National University Fungus Collection (SFC) and the National<br />

Institute <strong>of</strong> Biological Resources (NIBR) <strong>of</strong> the Ministry <strong>of</strong> Environment,<br />

a total <strong>of</strong> 523 decay fungi (104 species) associated with oak trees were identified<br />

and listed. No host specificity <strong>of</strong> decay fungi was evident. However, many decay<br />

fungi preferentially attacked specific sites within oak trees. To investigate the diversity<br />

<strong>of</strong> the ectomycorrhizal fungi, cloning and pyrosequencing were carried<br />

out. Through the sequence analysis <strong>of</strong> 243 clones, 67 sequences were identified<br />

as the plant and 176 sequences (62 phylotypes) were the fungi (Basidiomycota-<br />

142 (43), Ascomycota-34 (19)). Major phylotypes were ecotomycorrhizal fungi<br />

such as Russula and Lactarius, and the others were litter/decay rotters and dark<br />

septate endophytes. A total <strong>of</strong> 17,694 sequence reads were obtained from pyrosequencing,<br />

where 9,951 were fungi and 5,597 were plant. Compared with RFLP<br />

and cloning, the pyrosequencing approach yielded more diverse results.<br />

Little, Damon P. Cullman Program for Molecular Systematics, The New York<br />

Botanical Garden, Bronx, NY 10458. The use <strong>of</strong> DNA barcode techniques to<br />

identify the constituents <strong>of</strong> teas and herbal dietary supplements<br />

Dried fragmentary plant materials, such as those found in teas and herbal<br />

dietary supplements, are difficult to identify to species using only morphological<br />

characteristics. Fortunately, PCR amplifiable DNA can be extracted from most<br />

dried fragments and the resulting DNA barcodes can be used to reliably make<br />

identifications. For example, black cohosh (Actaea racemosa) herbal dietary supplements<br />

are commonly consumed to treat menopausal symptoms. Accidental<br />

misidentification and/or deliberate adulteration results in harvesting other, related,<br />

species that are then marketed as black cohosh. Some <strong>of</strong> these species are known<br />

to be toxic to humans. Two nucleotides in the plant barcode region consistently<br />

distinguish black cohosh from related species. Of 36 dietary supplements sequenced,<br />

27 (75%) have a sequence that exactly matches black cohosh. The remaining<br />

9 samples (25%) have a sequence identical to that <strong>of</strong> three Asian Actaea<br />

species (A. cimicifuga, A. dahurica, and A. simplex). Another example comes<br />

from commercial herbal tea products that are <strong>of</strong>ten composed <strong>of</strong> a number <strong>of</strong> plant<br />

species. Although barcode sequences could be generated for 60 <strong>of</strong> 71 (84%)<br />

herbal tea products examined, matching DNA identifications to listed ingredients<br />

was limited by incomplete databases, shared or nearly identical barcodes among<br />

some species, and a lack <strong>of</strong> standardized common names for all plant species.<br />

Analysis <strong>of</strong> the barcodes generated indicates that 25 (35%) <strong>of</strong> the herbal teas examined<br />

contained ingredients not mentioned on their labels.<br />

Lonergan, Erin R and Cathy L Cripps. Plant Sciences and Plant Pathology Department,<br />

Montana State University, Bozeman, MT 59715. The use <strong>of</strong> native ectomycorrhizal<br />

fungi in the restoration <strong>of</strong> whitebark pine<br />

Whitebark pine is a keystone species in high elevation ecosystems <strong>of</strong><br />

Western North <strong>America</strong>; it is rapidly declining throughout much <strong>of</strong> its range due<br />

to blister rust, a mountain pine beetle epidemic and fire suppression. The U.S. Fish<br />

and Wildlife Service found it warranted but currently precluded for protection<br />

under the Endangered Species Act. Huge restoration efforts are on going with<br />

over 200,000 nursery-grown seedlings planted in the Western U.S.; however,<br />

seedling survival rates are low. Ectomycorrhizal fungi can enhance conifer survival<br />

after out-planting, but previous to this research, whitebark pine seedlings<br />

were not inoculated before out-planting. Rhizopogon is <strong>of</strong>ten used for inoculation,<br />

but we found Suillus species specific to five needle pines to be important ectomycorrhizal<br />

partners for young and mature whitebark pine trees. Suillus species<br />

are also used for inoculation <strong>of</strong> cembran pines in Europe. Here the effects <strong>of</strong> inoculation<br />

<strong>of</strong> whitebark pine seedlings with native strains <strong>of</strong> Suillus sibiricus were<br />

investigated in greenhouse and field studies. In the greenhouse, lower levels <strong>of</strong><br />

fertilization resulted in a greater abundance <strong>of</strong> ectomycorrhizae, although all treatments<br />

with low N fertilizer allowed for colonization. Long containers promoted<br />

colonization, while short containers appeared to promote root disease. The field<br />

study in Waterton Lakes National Park, Alberta, Canada, is examining the survival<br />

<strong>of</strong> out-planted whitebark seedlings inoculated with native ectomycorrhizal<br />

fungi (Suillus sibiricus). A thousand seedlings were planted in “clusters” to mimic<br />

the planting strategy <strong>of</strong> Clark’s nutcracker. Each cluster included one, two, or<br />

three inoculated seedlings (in addition to non-inoculated controls). Clusters were<br />

planted in burned and un-burned areas, with and without beargrass. First year survival<br />

<strong>of</strong> seedlings was high with 95% <strong>of</strong> seedlings surviving regardless <strong>of</strong> treatment<br />

or habitat type. Monitoring in 2012 will examine survival rates in the second<br />

year when higher mortality rates are likely to occur.<br />

Lopez, Sara R 1 , Robert L Sinsabaugh 1 , David T Hanson 1 , and Andrea Porras-<br />

Alfaro 2 . 1 Department <strong>of</strong> Biology, University <strong>of</strong> New Mexico, Albuquerque, NM,<br />

87131, 2 Department <strong>of</strong> Biological Sciences, Western Illinois University, Macomb,<br />

IL, 61455. Estimating fungal contributions to nitrous oxide (N2O) emissions from nitrogen fertilized grasslands using tunable diode laser absorption<br />

spectroscopy.<br />

The use <strong>of</strong> nitrogen (N) fertilizers has created international concern as<br />

denitrification <strong>of</strong> fertilizer-derived nitrate could result in higher rates <strong>of</strong> N2O release<br />

to the atmosphere. The discovery that many eukaryotic fungi can transform<br />

N through nitrification and denitrification pathways is significant from both biochemical<br />

and ecological perspectives, because fungi dominate microbial metabolism<br />

in many terrestrial ecosystems. It is known that roots from semi-arid grasslands<br />

are colonized by a diverse fungal community, and suppression <strong>of</strong> such<br />

fungal biomass can cause a substantial reduction <strong>of</strong> N2O production. N2O production<br />

by six soil fungal species was estimated by analyzing gas samples from<br />

pure cultures isolated from blue-grama rhizosphere soils at the Sevilleta Long-<br />

Term Ecological Research site in Central New Mexico. Cultures were maintained<br />

in liquid medium containing nitrate as N source and incubated at 25°C for 10 days<br />

prior to sealing the flasks to limit aeration. Gas samples obtained after 24, 48 and<br />

72 hours <strong>of</strong> incubation were injected into a Tunable Diode Laser (Campbell Scientific).<br />

CO2 production was recorded to normalize N2O generation by fungal<br />

respiration. Preliminary results indicate that there is a high variability in N2O production<br />

among species but, overall, N2O production tends to increase with time<br />

<strong>of</strong> incubation. After 72 hours <strong>of</strong> incubation, the highest levels <strong>of</strong> N2O and CO2 were produced by Gibberella avenacea (4.<strong>63</strong>x10 -6 µmol 100 ml -1 mg dry mass -<br />

1 -1 -1<br />

and 0.3 µmol 100 ml mg dry mass respectively) and an unidentified group<br />

(Genbank AJ875391; 4.07x10 -4 µmol 100 ml -1 mg dry mass -1 and 4.95 µmol<br />

100 ml -1 mg dry mass -1 respectively.<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 29


Lueschow, Shiloh 1 , Lynnaun Johnson 1 , Robert McCleery 2 , Rod McClanahan 3 ,<br />

and Andrea Porras-Alfaro 1 . 1 Department <strong>of</strong> Biological Sciences, Western Illinois<br />

University, Macomb, IL, 2 Wildlife Ecology and Conservation, University <strong>of</strong><br />

Florida, Gainesville, FL, 3 USDA Forest Service, Shawnee National Forest, Harrisburg,<br />

IL. Morphological, physiological and molecular characterization <strong>of</strong><br />

different strains <strong>of</strong> Geomyces<br />

Geomyces destructans is a fungus responsible for the mycoses in bats<br />

known as White Nose Syndrome (WNS), which has already killed millions <strong>of</strong><br />

bats and may be driving some species towards extinction. Different Geomyces<br />

species are commonly found in the natural environment such as soils and caves,<br />

but it is not clear why G. destructans shows greater pathogenicity. G. destructans<br />

is considered a psychrophilic fungus (optimal grow is at low temperatures), but<br />

little is known about the growth requirements <strong>of</strong> other closely related Geomyces<br />

strains and species. The main goal <strong>of</strong> this study was to characterize the growth<br />

rates <strong>of</strong> various Geomyces isolates from Illinois and compare them to the<br />

pathogen G. destructans. Fourteen Geomyces isolates including G. destructans<br />

were plated on sabouraud dextrose agar. Isolates were identified using ITS rDNA.<br />

The isolates were obtained from bats collected in Illinois before the arrival <strong>of</strong><br />

WNS. Isolates were incubated at 25ºC and 6 ºC. Growth rates were measured<br />

weekly to determine differences between the various isolates and G. destructans.<br />

Geomyces destructans showed no growth after three weeks at 25º C. The other<br />

Geomyces isolates obtained from bats in Illinois showed growth after the first<br />

week at 25º C. In all cases except for G. destructans, the isolates grew faster at 25º<br />

C than at 6º C. Excluding G. destructans, the growth rate range for 25º C was 5<br />

mm/week to 9 mm/week. The growth rate range for 6º C was 3mm/week to<br />

5mm/week. Our study suggests that the psychrophilic (cold loving) nature <strong>of</strong> G.<br />

destructans may be an important factor influencing its pathogenecity.<br />

Luo, Jing and Ning Zhang. Department <strong>of</strong> Plant Biology and Pathology, Rutgers<br />

University, New Brunswick, New Jersey 08901. A phylogenetic study on the<br />

Magnaporthaceae (Ascomycota)<br />

The phylogenetic relationships among taxa in the Magnaporthaceae were<br />

investigated based on DNA sequences <strong>of</strong> multiple genes including ITS, SSU,<br />

LSU, RPB1, TEF and MCM7. The genera Magnaporthe and Gaeumannomyces<br />

were shown to be polyphyletic and their members were divided into four major<br />

groups based on the phylogenetic analyses. With the combination <strong>of</strong> morphological,<br />

biological and molecular data, we propose to establish and typify a new<br />

genus for the well-supported clade including M. poae, M. rhizophila, and G. incrustans.<br />

The current generic concepts <strong>of</strong> Gaeumannomyces, Magnaporthe and<br />

Pyricularia are revised. The constituent species <strong>of</strong> these genera are listed.<br />

Maltz, Mia R 1 , Kathleen Treseder 1 , and Jutta Burger 2 . 1 Department <strong>of</strong> Ecology<br />

and Evolutionary Biology, University <strong>of</strong> California, Irvine, CA, 2 Irvine Ranch<br />

Conservancy. Mycorrhizal associations in restored and invaded grasslands<br />

Mycorrhizal fungi perform a variety <strong>of</strong> ecosystem services, including facilitating<br />

plant establishment and improving nutrient uptake by plants. However,<br />

little is known about which Southern Californian mycorrhizal fungi facilitate native<br />

plant biodiversity in grasslands invaded by Brassica nigra. To compare<br />

among restoration sites, soil samples were taken from intact native grasslands<br />

(>50% native), manually managed B. nigra plots (30-50% native) highly invaded<br />

B. nigra cover (


microsatellite markers were developed and used to genotype ascospore progeny<br />

from 44 perithecia isolated from two nearly adjacent sites at West Rock Ridge<br />

State Park in New Haven, CT. The 13 markers ranged in allelic diversity from 2-<br />

15 alleles/locus, with an average <strong>of</strong> 4.3 alleles per locus in this population. Pooled<br />

sibling ascospore haplotypes for each <strong>of</strong> the 44 perithecia were used to infer 27<br />

unique multilocus diplotypes, from which parameters <strong>of</strong> population structure<br />

were estimated. Minimally significant (p≤0.05) linkage disequilibrium (LD) was<br />

observed between several pairs <strong>of</strong> loci, but a standardized estimate <strong>of</strong> multilocus<br />

LD was not significant. Genetic differentiation between the two subpopulations<br />

was minimally significant at three loci (Dest = 0.11, 0.17, 0.19), while the harmonic<br />

mean across all loci was not (Dest = 0.01); therefore, all subsequent analyses<br />

were performed on the entire population. Absence <strong>of</strong> segregation among sibling<br />

haplotypes was observed in 23 <strong>of</strong> the 44 perithecia, yielding a selfing rate, s,<br />

<strong>of</strong> 0.52. Heterozygote deficiencies were observed at all loci; Fis , the inbreeding<br />

coefficient, ranged from 0.39 to 1.0, and averaged 0.71 over all loci. Fis estimated<br />

from only the outcrossed portion <strong>of</strong> the population (n=21) was 0.42, supporting<br />

the hypothesis that biparental inbreeding contributed significantly to overall<br />

inbreeding in this population. These observations support theoretical models that<br />

posit the importance <strong>of</strong> biparental inbreeding to the evolutionary stability <strong>of</strong><br />

mixed mating.<br />

McCluskey, Kevin 1 and Jessie Glaeser 2 . 1 University <strong>of</strong> Missouri-KC, 5100<br />

Rockhill Road, Kansas City, MO 64110, 2 Center for Forest Mycology Research,<br />

USDA-Forest Service, NRS, Forest Products Laboratory, One Gifford Pinchot<br />

Drive, Madison, WI 53726. Research Coordination Network for US culture<br />

collections <strong>of</strong> microorganisms associated with plants<br />

A network bringing together scientists working with laboratory-based collections<br />

<strong>of</strong> microbes is being established. The network will hold numerous workshops<br />

teaching best practices for managing, preserving, and distributing bacteria, fungi,<br />

and other microscopic organisms in the context <strong>of</strong> formal culture collections.<br />

Biosecurity and regulatory issues will also be emphasized at workshops. Additional<br />

goals include re-establishing a pr<strong>of</strong>essional society <strong>of</strong> culture collection researchers<br />

in the US, developing internet based collection management tools, and<br />

fostering communication between US collections, foreign collections, and international<br />

collection networks. Collections <strong>of</strong> living microbes assure that current<br />

and past research and innovation are available to future generations <strong>of</strong> scientists.<br />

Biological materials that are made available via well-managed collections represent<br />

the foundation <strong>of</strong> the modern biotechnology industry. Materials in culture<br />

collections impact fields as diverse as human health, agricultural productivity,<br />

biotechnology, and biodiversity research. While there are several pr<strong>of</strong>essional collections<br />

in the US, most collections in the US are small and do not have long-term<br />

strategies for survival. The network supported by this grant will ensure that smaller<br />

collections benefit from expertise available at larger living microbe collections.<br />

NSF has recommended our RCN proposal for funding, but funding has not yet<br />

been approved. All interested parties are invited to participate.<br />

McCormick, Meghan A, Marc A Cubeta, and Larry F Grand. North Carolina<br />

State University, Department <strong>of</strong> Plant Pathology, Box 7567, Raleigh, NC 27695.<br />

Geographic distribution <strong>of</strong> Fomes fasciatus and F. fomentarius in the United<br />

States<br />

Fomes fasciatus and F. fomentarius can cause a white heart rot in multiple<br />

species <strong>of</strong> trees and play important roles in forest ecology and management.<br />

F. fasciatus is distributed across the Southeastern U.S., throughout Central <strong>America</strong><br />

and into South <strong>America</strong>., while F. fomentarius is distributed throughout the<br />

boreal forests <strong>of</strong> the Northern Hemisphere. North Carolina is a unique geographic<br />

transition area for these species as it represents the northern limit for F. fasciatus<br />

and the southern limit for F. fomentarius. In this study, U.S. distribution maps<br />

(by county) <strong>of</strong> F. fasciatus and F. fomentarius were developed based on records<br />

from 26 mycological herbaria, publications, and collections made for this study.<br />

The geographic distribution for both species was expanded to include six counties<br />

in four states not included in previous publications. Both species are associated<br />

with a diverse range <strong>of</strong> trees, which includes hardwood species from multiple genera.<br />

In the US, F. fasciatus has been reported predominantly on species <strong>of</strong> oak<br />

(Quercus) and hickory (Carya), while F. fomentarius typically decays species <strong>of</strong><br />

birch (Betula), beech (Fagus), and maple (Acer). The distribution <strong>of</strong> host species<br />

in many cases extends beyond the known limits <strong>of</strong> the distribution <strong>of</strong> F. fomentarius<br />

and F. fasciatus indicating that other delimiting factors contribute to the distribution<br />

and occurrence <strong>of</strong> these fungi.<br />

McCormick, Meghan A, Larry F Grand, and Marc A Cubeta. North Carolina<br />

State University, Department <strong>of</strong> Plant Pathology, Box 7567, Raleigh, NC 27695.<br />

Characterization <strong>of</strong> the wood decay fungi Fomes fasciatus and F. fomentarius<br />

using sequence analysis, morphology, and growth response to temperature<br />

in vitro<br />

The wood decay fungi Fomes fasciatus and F. fomentarius are endemic<br />

to the United States. Both fungi cause a white heart rot <strong>of</strong> multiple species <strong>of</strong> trees<br />

and have essential roles in forest ecology and management. Little is known about<br />

the genotypic and phenotypic diversity <strong>of</strong> these species. For this study, basidiocarps<br />

<strong>of</strong> F. fasciatus and F. fomentarius were collected from 13 states within the<br />

United States. Pure cultures <strong>of</strong> each species were isolated from basidiocarp context<br />

tissue and/or single basidiospores. Ten morphological characteristics were<br />

recorded for basidiocarps and pure cultures. F. fasciatus and F. fomentarius were<br />

morphologically similar, but could be differentiated based on the smaller average<br />

basidiospore size <strong>of</strong> F. fasciatus. The average colony diameter (mm) was measured<br />

every 3 d for 9 d using five isolates each <strong>of</strong> F. fasciatus and F. fomentarius<br />

on malt extract agar at temperatures ranging from 12 to 36 °C in 4 °C increments.<br />

The temperature optima <strong>of</strong> F. fasciatus and F. fomentarius were 32 °C and 28 °C,<br />

respectively. In a separate study, we found that F. fasciatus grew at 39 °C while<br />

F. fomentarius did not. The genetic relatedness <strong>of</strong> the two species was examined<br />

using maximum parsimony analysis <strong>of</strong> the internal transcribed spacer region<br />

(ITS) <strong>of</strong> the ribosomal DNA (rDNA) and the RNA polymerase II gene (RPB2).<br />

Preliminary results based on 100 replications indicate that the two species represent<br />

distinct lineages that form two distinct clades with 100 percent bootstrap support.<br />

Further analysis <strong>of</strong> ITS rDNA and RPB2 sequences will be discussed.<br />

McLaughlin, David J, TK Arun Kuman, and Rosanne Healy. Department <strong>of</strong><br />

Plant Biology, University <strong>of</strong> Minnesota, St. Paul, MN 55108. What can nuclear<br />

division, spindle pole body, and septal pore characters reveal about fungal<br />

phylogeny?<br />

A major goal <strong>of</strong> the AFTOL2 project has been to integrate molecular and<br />

structural data for key taxa at critical branches in the fungal tree <strong>of</strong> life to better<br />

understand phylogenetic relationships and character evolution. Nuclear division,<br />

spindle pole body (SPB) and septal pore characters have supported molecular<br />

phylogenies, but structural data is <strong>of</strong>ten absent for major clades or for taxa at key<br />

branch points within the fungal tree <strong>of</strong> life. We will consider recent data related to<br />

a number <strong>of</strong> hypotheses addressed by AFTOL2 and how the results <strong>of</strong> structural<br />

studies bear on them. These hypotheses include multiple losses <strong>of</strong> the flagellum<br />

and the transition to SPB’s in Fungi. In the Kickxellomycotina a ring-shaped SPB<br />

occurs in Coemansia reversa, which differs from SPB’s reported for other subphyla<br />

<strong>of</strong> zygomycetous fungi. The hypothesis that Orbiliomycetes is the earliest<br />

diverging lineage <strong>of</strong> Pezizomycotina gains support from septal pore analyses<br />

within the fruiting body <strong>of</strong> an Orbilia sp. and from ascus characters. The link between<br />

Ascomycota and Basidiomycota based on similarities in septal pore structure<br />

remains unresolved despite new insights into septal pore structure in Neolecta<br />

irregularis (Taphrinomycotina) and Helicogloea spp. (Pucciniomycotina),<br />

while a new septal type has emerged in Agaricomycotina through serial section<br />

reconstruction <strong>of</strong> the septal pore apparatus <strong>of</strong> Wallemia sebi. The latter supports<br />

the early divergence <strong>of</strong> the Wallemiomycetes. These data are retrievable for character<br />

analysis from the AFTOL Structural and Biochemical Database, but the<br />

database is a work in progress with a need for greater community involvement to<br />

make it more complete. Major gaps remain in the analysis <strong>of</strong> nuclear division,<br />

SPB and septal pore characters with some phyla and subphyla unstudied and a<br />

major phylum greatly understudied.<br />

McTaggart, Alistair R and M Catherine Aime. Department <strong>of</strong> Plant Pathology<br />

and Crop Physiology Louisiana State University Agricultural Center, 302 Life<br />

Sciences Bldg. Baton Rouge, LA 70803. ITS, problematic to amplify a rust<br />

barcode<br />

DNA barcoding <strong>of</strong>fers a rapid and accurate means to identify fungi and<br />

other organisms in comparison to traditional, morphological methods. Rusts (Pucciniales,<br />

Pucciniomycotina), the largest group <strong>of</strong> phytopathogenic fungi, are challenging<br />

for both morphological and molecular identification. They have a dynamic<br />

life cycle with five spore stages that infect different, <strong>of</strong>ten unrelated hosts.<br />

They are obligate parasites and cannot be maintained in pure culture, thus pure<br />

DNA is difficult, if not impossible to obtain. A molecular barcode that could differentiate<br />

between rust species and identify cryptic taxa is a necessity for diagnosticians<br />

and plant pathologists who rely mainly on morphology and host associations<br />

for species identification. One genus <strong>of</strong> rusts, Coleosporium, is<br />

taxonomically challenging and infects several agricultural and ornamental plants,<br />

such as pine and frangipani. Within northern <strong>America</strong> and Asia, C. asterum is believed<br />

to form a species complex. This study used Coleosporium as a model organism<br />

to investigate: 1. The systematic relationships <strong>of</strong> eight species <strong>of</strong><br />

Coleosporium in northern <strong>America</strong>, including the C. asterum species complex. 2.<br />

The utility <strong>of</strong> the ITS region as a molecular barcode for rust fungi. 3. The potential<br />

<strong>of</strong> other loci, such as the LSU region, as rust barcodes.<br />

Medina Rivera, Mariely and Matias J Cafaro. Department <strong>of</strong> Biology, University<br />

<strong>of</strong> Puerto Rico, Mayaguez, PR 00681. Micr<strong>of</strong>ungi community associated<br />

with yeast agriculture ant Cyphomyrmex minutus<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 31


Fungus-growing ants belong to the tribe Attini. They cultivate a single<br />

fungus (Basidiomycota) colony as food. To protect their cultivar from specific<br />

parasites in the genus Escovopsis (Ascomycota: Hypocreales) the ants engage in<br />

another association with antibiotic-producing Actinobacteria. In addition, the ants<br />

have specific hygienic behaviors that include farming and grooming <strong>of</strong> the cultivar.<br />

The ants also rearrange and transport the refused organic material inside and<br />

out <strong>of</strong> the nest. Because the cultivar pathogen is more abundant at the refuse material<br />

<strong>of</strong> the nest <strong>of</strong> other Attini ants in comparison with the cultivar garden is very<br />

important for the health <strong>of</strong> the cultivar to maintain this behavior. Cyphomyrmex<br />

minutus is the only Lower Attini species reported from Puerto Rico that practices<br />

yeast agriculture. We have investigated the micr<strong>of</strong>ungi community from both, the<br />

yeast cultivar and the refuse material piles. Our main objective was to detect and<br />

identify the cultivar pathogen Escovopsis and other micr<strong>of</strong>ungi associated with C.<br />

minutus. We obtained 152 cultures from 26 nests collected in Cambalache Forest<br />

in Puerto Rico. We also extracted total DNA from refuse samples and amplified<br />

the ITS region and cloned the products. The micr<strong>of</strong>ungi community associated to<br />

C. minutus is diverse. Three major classes <strong>of</strong> fungi are represented in our isolates<br />

(Ascomycota, Basidiomycota and Zygomycota). Aspergillus, Bionectria, Fusarium,<br />

Microdochium, Penicillium and Pestalotiopsis were shared between the cultivar<br />

and refuse material among 48 different genera detected. So far C. minutus<br />

yeast cultivar does not show signs <strong>of</strong> infection with Escovopsis or any other<br />

pathogens. We hypothesize that Escovopsis may not be present in Puerto Rico or<br />

that yeast agriculture is an effective adaptation to prevent pathogen infection.<br />

Mena-Ali, Jorge, Erica Goldberger, Elizabeth Heppenheimer, and Maggie Serpi.<br />

Department <strong>of</strong> Biology, Franklin & Marshall College, Lancaster PA 17603. An<br />

Evolutionary examination <strong>of</strong> ecological and physiological resistance within<br />

host-pathogen interactions: a Montiaceae and Microbotryum model<br />

Historically, coevolution is considered the primary process to justify disease<br />

evolution between hosts and their pathogens. However recent research has<br />

highlighted the relative importance <strong>of</strong> host shift events in the evolutionary divergence<br />

<strong>of</strong> pathogen species. The purpose <strong>of</strong> this project is to examine the evolutionary<br />

history <strong>of</strong> the association between smut fungi (Microbotryum spp.) that infect<br />

plants within the Montiaceae family. In this system, when fungal spores infect<br />

a susceptible host, a severe form <strong>of</strong> anther-smut disease develops, leading to complete<br />

sterilization <strong>of</strong> both male and female fertility in the plant. Through largescale<br />

herbarium surveys we were able to analyze the prevalence and global distribution<br />

<strong>of</strong> the disease. To date, we have surveyed 11,864 sheets from 21 herbaria<br />

representing 206 species. Of these, 51 sheets showed disease giving a 0.43% disease<br />

rate. However, if we exclude those species without signs <strong>of</strong> disease, the disease<br />

rate increases to 9.9%. Additionally, we collected spore samples from infected<br />

specimens for molecular analysis. These samples were analyzed with<br />

fungal-specific markers (NADH, ITS). Initial phylogenetic analysis suggests a<br />

pattern <strong>of</strong> divergence among Microbotryum isolates that reflects host taxonomic<br />

classification. However, geographic distribution also seems to have contributed to<br />

isolation and speciation. These results will be used to analyze the evolutionary history<br />

<strong>of</strong> Microbotryum species associated with Montiaceae in the context <strong>of</strong> varying<br />

pathogenicity and host specificity.<br />

Menolli, Nelson, Jr 1,2 and Marina Capelari 1 . 1 Instituto de Botânica, Núcleo de<br />

Pesquisa<br />

2<br />

em Micologia, Caixa Postal 68041, 04045-972 São Paulo, SP, Brazil,<br />

Instituto Federal de Educação, Ciência e Tecnologia de São Paulo, Campus São<br />

Paulo, CCT / Biologia, Rua Pedro Vicente 625, 01109-010 São Paulo, SP, Brazil.<br />

One hundred and twelve years <strong>of</strong> Pluteus knowledge in Brazil: revision <strong>of</strong> the<br />

first collections studied by J. Rick and P. Hennings<br />

The genus Pluteus comprises ca. 300 species worldwide <strong>of</strong> which 70 have<br />

been mentioned from Brazil; however, it is believed that about 23 are certainly<br />

known. To solve some <strong>of</strong> the taxonomic problems, techniques <strong>of</strong> molecular biology<br />

can be used, but there is some limitations e.g. when studying old collections,<br />

and an accurate morphological study may be an alternative. Complementing the<br />

Brazilin knowledge <strong>of</strong> Pluteus, the first collections studied in the early 20th century<br />

were re-examined. At first, Hennings described P. scruposus, P. cervinus var.<br />

griseoviridis and P. termitum. Later, between 1907 and 1961, Rick recorded 21<br />

more Pluteus taxa. The holotypes <strong>of</strong> P. scruposus and P. termitum deposited in<br />

Berlin were probably destroyed. Collections <strong>of</strong> nine taxa recorded by Rick (P.<br />

cervinus var. patricius, P. cristatulus Rick, P. eximius, P. fibrillosus Rick, P.<br />

melanodon, P. nanus var. podospileus, P. umbrosus) also were not found. All<br />

other collections are deposited at PACA and SP and despite the bad conservation,<br />

some micromorphological structures were recovered and the following considerations<br />

could be made correcting the European names attributed in the past by Rick<br />

and Hennings. P. brunneopictus: is probably P. tucumanus Singer; P. cervinus<br />

and P. cervinus var. griseoviridis: represent P. xylophillus (Speg.) Singer; P. exiguus:<br />

is probably P. jamaicensis Murrill; P. granulatus: is possibly P. glaucotinctus<br />

Horak; P. hispidulus: is probably P. yungensis Singer; P. leptonia Rick: is<br />

a species <strong>of</strong> Entoloma s.s.; P. nanus: is possibly P. pulverulentus Murrill; P. pellitus:<br />

is probably P. petasatus; P. phlebophorus: represents two collections which<br />

32 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

probably are P. pulverulentus and P. tucumanus; P. sensitivus Rick: is probably<br />

P. albostipitatus; P. straminellus Rick: is a nomen dubium; P. velatus Rick: is a<br />

nomen dubium; P. wehlianus: is a species <strong>of</strong> Bolbitiaceae or Strophariaceae.<br />

Methven, Andrew S 1 and Andrew N Miller 2 . 1 Department <strong>of</strong> Biological Sciences,<br />

Eastern Illinois University, Charleston, IL 61920, 2 Illinois Natural History<br />

Survey, University <strong>of</strong> Illinois, Champaign, IL 61820. Evolutionary relationships<br />

<strong>of</strong> the gomphoid genus Clavariadelphus: One genus or two?<br />

The genus Clavariadelphus includes a group <strong>of</strong> club-shaped basidiomes<br />

most commonly collected in late summer and fall in northern, boreal forests<br />

throughout North <strong>America</strong>. A monograph <strong>of</strong> Clavariadelphus in North <strong>America</strong><br />

divided the genus into two subgenera: subgenus Clavariadelphus which includes<br />

ectomycorrhizal species that are associated with coniferous or deciduous trees,<br />

broadly ellipsoid basidiospores (length:width ratio < 2.5), and little or no hyphae<br />

at the base <strong>of</strong> the basidiomes; and, subgenus Ligulus with saprotrophic species<br />

that function ecologically as litter decomposers in coniferous forests, narrowly ellipsoid<br />

basidiospores (length:width ratio > 2.5), and copious amounts <strong>of</strong> hyphae<br />

which bind the substrate to the base <strong>of</strong> the basidiomes. While some agaricologists<br />

have argued that these two subgenera are distinct enough to be recognized as separate<br />

genera, questions about the range <strong>of</strong> variation in morphological characters,<br />

chemical spot tests and cultural characters have precluded recognition <strong>of</strong> the two<br />

groups as segregate genera. We hypothesize that the genus Clavariadelphus is<br />

polyphyletic and, in order to adhere to a natural system <strong>of</strong> classification, needs to<br />

be subdivided into two monophyletic groups or genera. Two nuclear ribosomal<br />

genes, ITS and LSU were amplified, sequenced and analyzed in a phylogenetic<br />

context to determine if Clavariadelphus should be segregated into two genera.<br />

Based on ITS and LSU sequences, Subgenus Ligulus is well supported as a monophyletic<br />

group that is distinct from and basal to Subgenus Clavariadelphus. Since<br />

the type species <strong>of</strong> the genus, C. pistillaris, belongs to Subgenus Clavariadelphus,<br />

a new genus will be proposed for the taxa included in Subgenus Ligulus.<br />

Miadlikowska, Jolanta 1 , Bernie Ball 1 , Francesc López-Giráldez 2 , Jeffrey P<br />

Townsend 2 , Ester Gaya 1 , Tami McDonald 1 , Suzanne Joneson 1 , Andrii Gryganskyi<br />

1 , Teresita M Porter 1 , Brandon Matheny 3 , Kassian Kobert 4 , Alexandros Stamatakis<br />

4 , Barbara Robbertse 5 , Joseph Spatafora 5 , David Hibbett 6 , Rytas Vilgalys<br />

1 , and François Lutzoni 1 . 1 Department <strong>of</strong> Biology, Duke University,<br />

Durham, NC 27708, 2 Department <strong>of</strong> Ecology and Evolutionary Biology, Yale<br />

University, New Haven, CT 06520, 3 Department <strong>of</strong> Ecology and Evolutionary<br />

Biology, University <strong>of</strong> Tennessee, Knoxville, TN 37996, 4 Heidelberg Institute for<br />

Theoretical Studies, Schloss-Wolfsbrunnenweg 35, D-69118 Heidelberg, Germany,<br />

5 Department <strong>of</strong> Botany and Plant Pathology, Oregon State University,<br />

Corvallis, OR 97331, 6 Department <strong>of</strong> Biology, Clark University, Worcester, MA<br />

01610. Novel molecular markers and their utility in molecular systematics <strong>of</strong><br />

Fungi<br />

Although next generation sequencing methods have proven to be very<br />

successful in accelerating data acquisition, selecting the optimal set <strong>of</strong> molecular<br />

markers for phylogenetic studies has remained a complex endeavor. Existing fungal<br />

phylogenies demonstrate the urgent need for novel single-copy protein-coding<br />

genes to resolve phylogenetic relationships among fungi at all taxonomic levels<br />

with high confidence. As part <strong>of</strong> the Assembling the Fungal Tree <strong>of</strong> Life project<br />

(AFToL 2), a comparative genomic approach was adopted to select all singlecopy<br />

orthologous genes with the greatest potential to resolve the most challenging<br />

supraordinal nodes <strong>of</strong> the fungal tree <strong>of</strong> life. Based on the comparison <strong>of</strong> 39<br />

fungal genomes, 71 potentially single-copy orthologous genes were selected, and<br />

a total <strong>of</strong> 243 universal primer pairs were designed and tested on six exemplar<br />

species representing Ascomycota, Basidiomycota and early-diverging fungi. The<br />

successful amplification and sequencing <strong>of</strong> 19 new gene regions adds a total <strong>of</strong><br />

ca. 13,000 bp per taxon, in addition to the commonly sequenced nucLSU, nuc-<br />

SSU, mitSSU, MCM7, RPB1 and RPB2. The final datasets used to evaluate the<br />

performance <strong>of</strong> these new genes include 37 non-lichenized fungi, for which genomic<br />

sequence data are available, and up to 24 lichen-forming members <strong>of</strong> the<br />

Arthoniomycetes, Dothideomycetes, Eurotiomycetes, Lecanoromycetes, and<br />

Lichinomycetes, for which the sequences were obtained mostly from cultures <strong>of</strong><br />

the mycobionts using single-gene Sanger sequencing, but also through next generation<br />

genome sequencing. Maximum likelihood analyses were completed on<br />

each <strong>of</strong> the 19 novel and 8 commonly used gene regions independently and on<br />

various multi-locus combinations on the same set <strong>of</strong> taxa when possible. Phylogenetic<br />

efficiency (i.e., the level <strong>of</strong> resolution and internode robustness) and phylogenetic<br />

informativeness (sensu Townsend) among all loci was compared. The<br />

performance <strong>of</strong> novel versus older genes used to infer phylogenetic relationships<br />

among fungi is also compared.<br />

Miller, Stephen J, Jr 1 , Hayato Masuya 2 , and Ning Zhang 1 . 1 Dept Plant Biology<br />

and Pathology, Rutgers University, New Brunswick, NJ 08901, USA, 2 Dept<br />

Forest Microbiology, Forestry & Forest Products Research Institute, Matsunosato<br />

Continued on following page


1, Tsukuba, Ibaraki 305-8687, JAPAN. Assessing the effects <strong>of</strong> radiation after<br />

the Fukushima Daiichi meltdown on dogwood endophyte communities<br />

Radionuclide contamination in the environment is one <strong>of</strong> the predominate<br />

concerns to human and environmental health. Radionuclide contamination enters<br />

the environment in the form <strong>of</strong> fallout from nuclear weapons, nuclear waste generated<br />

from industries, medical use <strong>of</strong> radioisotopes, and accidents from nuclear<br />

meltdowns. Previously, the role <strong>of</strong> fungi played in mediating radionuclide movement<br />

after the Chernobyl accident was examined. These studies were conducted<br />

a few years after the initial accident. Recently, the Fukushima Daiichi nuclear<br />

meltdown has released radionuclides in the environment, with radiation levels detectable<br />

in North <strong>America</strong>. In this study, we analyzed the fungal endophytic communities<br />

associated with Cornus kousa (Kousa dogwood) leaf samples collected<br />

in Japan in June 2010 and June 2011, which was three months after the initial incident.<br />

For the June 2010 samples, 320 <strong>of</strong> 600 leaf segments from 6 samples<br />

yielded fungal cultures. For the June 2011 samples, fungi were recovered in 116<br />

<strong>of</strong> 400 leaf segments from 4 samples located 60-100 km from the Fukushima Daiichi<br />

site. In the June 2010 samples, Colletotrichum and Penicillium were the dominant<br />

genera. Cosmospora, Phomopsis, and Diaporthe were the dominant genera<br />

in the samples recovered after the nuclear meltdown. More melanized fungi were<br />

isolated from the 2011 dogwood samples than those collected in 2010. The results<br />

indicate that the sampled fungal endophyte communities changed significantly<br />

from 2010 to 2011, likely affected by radiation pollution.<br />

Minnis, Andrew M. Center for Forest Mycology Research, Northern Research<br />

Station, USDA-Forest Service, Madison, WI, 53726. Ascomycete taxonomy<br />

and the outset <strong>of</strong> the one name era for fungi<br />

Multiple correct names, including one for each stage <strong>of</strong> non-lichen forming<br />

ascomycetes with pleomorphic life cycles, have been allowed under formalized<br />

Codes used to name fungi for a significant period <strong>of</strong> the history <strong>of</strong> mycology.<br />

Due to philosophical preferences <strong>of</strong> a large group <strong>of</strong> mycologists, the most<br />

important article that allowed for multiple names (Art. 59) was essentially deleted<br />

at the 2011 Melbourne International Botanical Congress. The replacement text<br />

(Art. 59) in the International Code <strong>of</strong> Nomenclature for algae, fungi, and plants allows<br />

only one correct name for non-lichen forming ascomycetes with pleomorphic<br />

life cycles. Although already enacted, these changes, which will formally and<br />

retroactively take effect in 2013 and have substantial impact on how mycologists<br />

and other users <strong>of</strong> fungal names will communicate about fungi, will be summarized.<br />

Pragmatic considerations will be noted and illustrated by examples <strong>of</strong> the<br />

impact on ascomycete taxonomy at the various ranks, especially in regards to priority,<br />

legitimacy, and validity <strong>of</strong> names as well as author citations and types. Some<br />

observations on what in general will be “the rise <strong>of</strong> the anamorphs” will be made.<br />

The means by which the mycological community can avoid disadvantageous<br />

changes to names during the transitional period to one name while maintaining a<br />

connection to important historical works will be outlined and a call for much<br />

needed participation and help in the process will be shared. The importance <strong>of</strong><br />

studying the whole fungus and integrating the taxonomy <strong>of</strong> anamorphs and teleomorphs<br />

in future works will be emphasized.<br />

Mohan, Jacqueline E 1 , Jerry M Melillo 2 , James S Clark 3 , Fahkri A Bazzaz 4 ,<br />

Katherine J Bridges 1 , Charles Cowden 1 , Paul Frankson 1 , Richard Shefferson 1 ,<br />

Richard Sicher 5 , Timothy Sipe 6 , and Lewis Ziska 5 . 1 Odum School <strong>of</strong> Ecology,<br />

University <strong>of</strong> Georgia Athens, GA 30602, 2 The Ecosystems Center, Marine Biological<br />

Laboratory Woods Hole, MA 02543, 3 Nicholas School <strong>of</strong> the Environment<br />

Duke University Durham, NS 27278, 4 Harvard University Department <strong>of</strong><br />

Organismic and Evolutionary Biology Cambridge, MA 02138, 5 USDA, Crop<br />

Systems and Global Change Beltsville, MD 20705-2350, 6 Franklin and Marshall<br />

College Lancaster, PA 17604-3003. The arbuscular advantage: Soil warming<br />

stimulates the growth <strong>of</strong> arbuscular mycorrhizal (AM)-associated tree<br />

species and poison ivy over 8 years at Harvard Forest<br />

Impacts <strong>of</strong> warmer soils for nutrient biogeochemical cycling are fairly<br />

well known, particularly from soil warming experiments situated on fertile soils<br />

<strong>of</strong> mid-high latitudes; namely, warmer soils cycle nitrogen (N), carbon (C), and<br />

in one previous study, phosphorus (P) faster. Potential impacts <strong>of</strong> warmer soils<br />

and increased availabilities <strong>of</strong> inorganic soil nutrients on plant-mycorrhizal interactions<br />

are less clear, as is the scaling up <strong>of</strong> plant-mycorrhizal relations with<br />

warming for ecosystem composition and functioning. Here we present results<br />

from the Harvard Forest large-plot (900-m 2 ) soil warming experiment demonstrating<br />

that warmer soils with higher availabilities <strong>of</strong> inorganic N enhance the<br />

growth <strong>of</strong> arbuscular mycorrhizal (AM)-associated woody species with important<br />

implications for future carbon sequestration by forests. In the pre-treatment growing<br />

seasons <strong>of</strong> 2001 and 2002, juvenile trees had similar growth rates in the to-be-<br />

Heated and the Control plots. Following warming, commenced May 2003, juvenile<br />

trees associating with AM fungi displayed significant and consistent growth<br />

increases in the Heated plot. This enhanced growth was sustained over time<br />

(2003-2010) and by the following taxa: sugar maple (Acer saccharum), red maple<br />

(A. rubrum), black cherry (Prunus serotina), white ash (Fraxinus americana), and<br />

poison ivy (Toxicodendron radicans). Co-dominant oaks, pines, and birches did<br />

not grow faster with warming. Further, mycorrhizal root colonization was significantly<br />

less in the maple species growing under warming while foliar %N and %P<br />

values were maintained, suggesting faster growth <strong>of</strong> AM-species may be due to<br />

decreased C costs <strong>of</strong> mycorrhizal associates. These results suggest a future shift<br />

in eastern temperate composition towards less productive tree species with important<br />

implications for future carbon dynamics. These finding may partially explain<br />

the increasing abundance <strong>of</strong> red maple and woody vines observed over the<br />

last several decades in eastern forests and globally, respectively.<br />

Moll, Julia 1,2 , Dörte Dibbern 3 , Susanne Kramer 4 , François Buscot 1,2 , Dirk<br />

Krüger 1 , Tillmann Lueders 3 , Sven Marhan 4 , and Ellen Kandeler 4 . 1 UFZ -<br />

Helmholtz Centre for Environmental Research, Department <strong>of</strong> Soil Ecology, Th.-<br />

Lieser-Str. 4, D-06120 Halle/Saale, Germany, 2 University <strong>of</strong> Leipzig, Institute <strong>of</strong><br />

Biology I, Johannisallee 21-23, D-04103 Leipzig, Germany, 3 Helmholtz Zentrum<br />

München, German Research Center for Environmental Health, Institute <strong>of</strong><br />

Groundwater<br />

4<br />

Ecology, Ingolstädter Landstr. 1, D-85764 Neuherberg, Germany,<br />

University <strong>of</strong> Hohenheim, Institute <strong>of</strong> Soil Science and Land Evaluation, Soil<br />

Biology Group, D-70593 Stuttgart, Germany. Fungal and bacterial diversity in<br />

connection with soil decomposition processes resolved using 13 C rRNA-SIP<br />

Soil fungi and bacteria play an important role in global nutrient cycles because<br />

<strong>of</strong> their decomposition <strong>of</strong> organic materials. The use <strong>of</strong> rRNA-based stable<br />

isotope probing (SIP) allows us to follow the carbon flux into soil microbial communities.<br />

Using this novel approach, we aimed at quantifying the carbon flow<br />

from plant-derived substrates into the bacterial and fungal channel at different<br />

stages <strong>of</strong> decomposition and to link these microorganisms to specific degradation<br />

processes. Highly 13 C-labeled labile (glucose) and recalcitrant (cellulose, maize<br />

roots, maize leaves) substrates were mixed into soil microcosms and incubated for<br />

32 days. Following cesium chloride density gradient ultracentrifugation <strong>of</strong> RNA<br />

extracts<br />

13<br />

(separating heavy labeled from light unlabeled RNA) from controls and<br />

C-microcosms sampled at 2, 8, 16 and 32 days post-incubation, rRNA was converted<br />

to cDNA and amplified. T-RFLP on the SSU and 16S amplicons was applied<br />

to detect the fungi and bacteria utilizing these substrates. It was possible to<br />

identify carbon-assimilating microbes for all substrates and time points representing<br />

substrate degradation stages. For both fungi and bacteria, the community (<strong>of</strong><br />

operational taxonomic units, OTUs) that consumed carbon from labile substrate<br />

clearly differed from that consuming it from recalcitrant material. The bacterial<br />

communities that appeared to be actively assimilating carbon in an early stage <strong>of</strong><br />

decomposition differed from that active in later stages. The fungal utilizers, however,<br />

rather remained the same over time. To validate our results and to better<br />

identify active taxa, clone library or pyrosequencing approaches are currently<br />

added onto the analysis <strong>of</strong> our experiment. By using SIP, this study will thus gain<br />

first insights into the relationship between carbon quality and availability and it<br />

will link soil microbial taxonomic with functional diversity.<br />

Monacell, James T and Ignazio Carbone * . Department <strong>of</strong> Plant Pathology,<br />

North Carolina State University, Raleigh, NC 27606. Mobyle SNAP Workbench:<br />

a phylogenetic and genomic analysis web portal<br />

The SNAP Workbench toolkit is a stand-alone s<strong>of</strong>tware application that integrates<br />

a wide array <strong>of</strong> bioinformatics tools for phylogenetic and population genetic<br />

analyses. We have developed a web-portal frontend, using the Mobyle portal<br />

framework, which runs all <strong>of</strong> the programs available in the stand-alone SNAP<br />

Workbench on a high-performance Linux cluster. Additionally, we have expanded<br />

the selection <strong>of</strong> programs to over fifty tools, including 1) a suite <strong>of</strong> genome assembly<br />

and analysis tools such as Fastx-Toolkit, Quake, Burrows-Wheeler Alignment<br />

Tool, Velvet, SOAPdenovo, GeneMark-ES, and Augustus, 2) large-scale<br />

phylogenetic methods as implemented in RAxML, and 3) metagenomic analysis<br />

tools such as Mothur and Esprit. The Mobyle SNAP Workbench web-portal allows<br />

researchers to seamlessly execute and manage otherwise complex commandline<br />

programs with multiple input files and parameters, as exemplified in the coalescent<br />

approaches implemented in Genetree, MDIV, IM, IMa, MIGRATE, and<br />

LAMARC; these tools are parameter-rich and therefore computationally intensive,<br />

<strong>of</strong>ten requiring several days or weeks to run. Our optimization includes 1) selecting<br />

appropriate number <strong>of</strong> compute nodes and machine architecture for MPI programs<br />

and 2) parallelization <strong>of</strong> computational tasks to execute on multiple machines<br />

when possible (e.g. Genetree). A unique feature <strong>of</strong> the portal is the<br />

implementation <strong>of</strong> workflows that link together several programs sequentially.<br />

This greatly facilitates exploratory analyses for the beginner but also allows for efficient<br />

use <strong>of</strong> cluster resources. The possibility <strong>of</strong> making the portal accessible to<br />

the public, as a web service on Amazon EC2 or at NC State, will be discussed.<br />

Moore, Geromy G and Kenneth C Ehrlich. Southern Regional Research Center,<br />

USDA, ARS, New Orleans, LA, USA. Competition experiments to test the efficacy<br />

<strong>of</strong> Aspergillus flavus biocontrol strains<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 33


Several non-aflatoxigenic strains <strong>of</strong> Aspergillus flavus have been in use<br />

for nearly a decade to inhibit colonization <strong>of</strong> important food commodities by toxigenic<br />

Aspergilli such as A. flavus and A. parasiticus. Annual application <strong>of</strong> these<br />

biocontrol strains is necessary because they do not appear to persist in the field.<br />

The reason(s) for this are unknown. One possible mechanism to explain the loss<br />

<strong>of</strong> an atoxigenic biocontrol competitor strain is that it could be recombining with<br />

the native toxigenic population and reacquiring toxigenic properties. A second<br />

possibility is that aflatoxin production increases the likelihood <strong>of</strong> survival for toxigenic<br />

strains in the soil, during overwintering, compared to that <strong>of</strong> an atoxigenic<br />

strain. To test the relative efficacy <strong>of</strong> these non-aflatoxigenic strains as bio-competitors,<br />

and to assess the reason for the loss <strong>of</strong> the field inoculum with time (posttreatment),<br />

we transformed six different biocontrol strains with a plasmid containing<br />

DNA for expressing an ‘enhanced green fluorescent protein’ (eGFP) tag.<br />

Competition experiments in petri plates have been set up to test the efficacy <strong>of</strong> the<br />

transformed biocontrol strains to colonize cotton seed. Tests included: GFP-transformed<br />

biocontrol strains vs non-transformed (homologous and heterologous)<br />

biocontrol strains, and GFP-transformed biocontrol strains vs toxigenic strains <strong>of</strong><br />

A. flavus and A. parasiticus. These in vitro studies will test if the GFP tag diminishes<br />

the aggressiveness <strong>of</strong> the biocontrol strain, as well as the relative abilities <strong>of</strong><br />

the individual strains to compete against toxigenic strains. If these tests show<br />

promise in using GFP-tagged strains to track the introduced fungi, subsequent<br />

studies will be done under conditions imitating the growing environment <strong>of</strong> the<br />

cotton plant in order to track the longevity and recombining potential <strong>of</strong> these biocontrol<br />

strains.<br />

Morgado, Luis N 1 , Manon Neilen 1 , Machiel E Noordeloos 1 , D Lee Taylor 2 , Ina<br />

Timling 2 , and József Geml 1 . 1 Netherlands Centre for Biodiversity Naturalis, Leiden<br />

University, P.O. Box 9514, Einsteinweg 2, 2300 RA Leiden, The Netherlands,<br />

2 Institute <strong>of</strong> Arctic Biology, University <strong>of</strong> Alaska Fairbanks, Fairbanks,<br />

AK 99775-7000. Phylogenetic diversity <strong>of</strong> the ectomycorrhizal genus Cortinarius<br />

(Agaricales, Basidiomycota) in the Arctic<br />

Mycorrhizal associations are abundant and widespread in almost all<br />

ecosystems, and c. 80% <strong>of</strong> land plant species form associations with mycorrhizal<br />

fungi. They play a particularly important role in the functioning <strong>of</strong> terrestrial arctic<br />

ecosystems, where arctic plants are highly dependent on mutualistic relationships<br />

with mycorrhizal fungi for survival in these nutrient-poor environments. Ectomycorrhiza<br />

(ECM) is the predominant mycorrhiza type in arctic and alpine<br />

environments, and ECM fungi are crucial for the survival <strong>of</strong> arctic shrubs (e.g. Betula,<br />

Dryas, Salix). Although recent molecular studies have revealed high diversity<br />

in arctic ECM communities, the systematic treatment <strong>of</strong> several arctic ECM<br />

genera is still not adequate. Cortinarius is one <strong>of</strong> the most abundant genera in the<br />

Arctic. Here we present preliminary results on the phylogenetic diversity <strong>of</strong> the<br />

genus Cortinarius in the North <strong>America</strong>n and European Arctic. We analyzed internal<br />

transcribed spacer (ITS) rDNA sequences from basidiomata and soil samples<br />

collected in northern Alaska and Svalbard using likelihood-based phylogenetic<br />

methods. We detected at least 28 phylogroups, from which 18 matched<br />

sequences from sporocarps deposited in public databases. These included C. umbilicatus,<br />

C. biformis, C. delibutus, C. favrei, C. cinnamomeus, C. aureomarginatus,<br />

C. urbicus, C. fulvescens, among others. The other 8 represent previously<br />

unsequenced taxa that may or may not be newly discovered species. Future<br />

investigations will include multi-gene phylogenetics and morphological analyses.<br />

Morgado, Luis N 1 , Machiel E Noordeloos 1 , Delia Co-David 1 , Yves Lamourex<br />

2 , and József Geml 1 . 1 National Herbarium <strong>of</strong> the Netherlands, Netherlands<br />

Centre for Biodiversity Naturalis, Leiden University, P.O. Box 9514, Einsteinweg<br />

2, 2300 RA Leiden, The Netherlands, 2 505, Rue Saint-Alexandre, app. 401,<br />

Longueuil (Québec), Canada, J4H3G3. Biogeographic and phylogenetic relationships<br />

<strong>of</strong> four easily recognizable morphospecies <strong>of</strong> Entoloma Section Entoloma<br />

(Basidiomycota), inferred from molecular and morphological data<br />

Species from Entoloma section Entoloma are commonly recorded from<br />

both the Northern and Southern Hemispheres and, according to literature, most <strong>of</strong><br />

them have, at least, Nearctic-Palearctic distribution. However, all records are<br />

based on morphological analysis, and studies relating morphology, molecular data<br />

and species distribution are lacking. In this study, we selected four morphospecies<br />

from Section Entoloma, to answer specific questions considering species concept<br />

and geographic distribution: E. sinuatum (E. lividum auct.), E. prunuloides (typespecies<br />

<strong>of</strong> section Entoloma), E. nitidum and the European red-listed E. bloxamii,<br />

with collections from Europe, North <strong>America</strong>, Australia and New Zealand. We<br />

combined molecular phylogenetics (based on nuclear LSU, ITS and rpb2 and mitochondrial<br />

SSU sequences) with morphological analysis to infer interspecific relationships<br />

and distribution patterns. Our results indicate that most species appear<br />

to have more restricted distribution than previously assumed. None <strong>of</strong> the collections<br />

studied from North Hemisphere and South Hemisphere proved to be conspecific.<br />

Entoloma sinuatum and E. nitidum contain phylogeographical groups<br />

that are partly recognizable using morphological characters. Entoloma bloxamii is<br />

paraphyletic, falling apart into 3 distinct lineages, one represented in Europe, a<br />

34 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

second restricted to North <strong>America</strong>, and a third represented in both these continents.<br />

The European and North <strong>America</strong>n E. prunuloides appear to be conspecific.<br />

Furthermore the accepted taxa appear to belong to two distinct clades, (1) Entoloma<br />

clade, with E. sinuatum, E. subsinuatum, E. whiteae, E. flavifolium, and<br />

(2) Prunuloides clade, with E. prunuloides, E. bloxamii and E. nitidum.<br />

Morgenstern, Ingo 1,2 , Justin Powlowski 1,3 , and Adrian Tsang 1,2 . 1 Centre for<br />

Structural and Functional Genomics, Concordia University, 7141 Sherbrooke<br />

Street West, Montreal, Quebec, Canada, 2 Department <strong>of</strong> Biology, Concordia<br />

University, 3 Department <strong>of</strong> Chemistry and Biochemistry, Concordia University.<br />

Transcriptional activity <strong>of</strong> “GH61”-encoding genes from various fungal<br />

species grown on straws<br />

The efficient degradation <strong>of</strong> lignocellulosic biomass using microbial enzymes<br />

is considered a key step for the production <strong>of</strong> second generation bi<strong>of</strong>uel<br />

and value-added by-products. Efforts to improve commercial cellulase mixtures<br />

include the spiking with auxiliary enzymes. Recently, members <strong>of</strong> the so-called<br />

glycosyl hydrolase family 61 (GH61) have attracted much interest due to their cellulase-enhancing<br />

properties in enzyme mixtures. Copies for GH61 encoding<br />

genes are present in the majority <strong>of</strong> fungal genomes; however, the copy number<br />

<strong>of</strong> GH61s differs tremendously among the investigated genomes, exceeding thirty<br />

copies in some species. Currently, it is not known whether this is merely an example<br />

<strong>of</strong> high redundancy or how far functional differences are present among<br />

GH61 paralogs. We are examining the transcriptional pr<strong>of</strong>ile <strong>of</strong> “GH61s” both<br />

from ascomycete and basidiomycete species grown on alfalfa and/or barley straw<br />

as substrate and are interpreting the results in a phylogenetic context.<br />

Morrison, Eric W 1 , Serita D Frey 2 , W Kelley Thomas 3 , and Anne Pringle 4 1<br />

.<br />

Department <strong>of</strong> Molecular, Cellular and Biomedical Sciences, University <strong>of</strong> New<br />

Hampshire, Durham, NH, 2 Department <strong>of</strong> Natural Resources and the Environment,<br />

University <strong>of</strong> New Hampshire, Durham, NH, 3 Hubbard Center for Genome<br />

Studies, University <strong>of</strong> New Hampshire, Durham, NH, 4 Organismic and Evolutionary<br />

Biology, Harvard University, Cambridge, MA. Diversity and composition<br />

<strong>of</strong> soil fungal communities under long-term nitrogen enrichment<br />

Nitrogen (N) deposition from fossil fuel burning has the potential to affect<br />

ecosystem processes such as the decomposition and storage <strong>of</strong> soil organic matter.<br />

The Harvard Forest Chronic Nitrogen Addition experiment (HFCN) was established<br />

in 1989 to test the effects <strong>of</strong> long-term N fertilization on ecosystem<br />

processes in a northeastern mixed-hardwood forest. Three plots receive one <strong>of</strong><br />

three treatments: ambient N deposition (control), 50 kg N ha -1 yr -1 (low N), or<br />

150 kg N ha -1 yr -1 (high N). Researchers at this site have observed an accumulation<br />

<strong>of</strong> soil C in the N fertilized plots and a decrease in fungal biomass, ligninolytic<br />

enzyme activity, and rates <strong>of</strong> litter decay. Soil fungi are the primary decomposers<br />

<strong>of</strong> lignin in these communities. We hypothesized that decreased<br />

decomposition rates in N fertilized plots may be due to decreased diversity and<br />

changes in the composition <strong>of</strong> the fungal community. We performed a marker<br />

gene study <strong>of</strong> the fungal community in the organic soil horizon using 454 sequencing<br />

<strong>of</strong> three loci: ITS1, ITS2, and rDNA large subunit D2-D3 region. The<br />

dominant fungal family in soils under ambient N deposition was the Russulaceae.<br />

This family underwent a significant decrease in relative abundance in N treated<br />

soils. Unknown fungi dominated N treated soils. Control soils had significantly<br />

fewer unknown OTUs. High N soils had higher numbers <strong>of</strong> OTUs and singleton<br />

OTUs then control soils and low N soils, and had higher predicted richness. Fungal<br />

communities in high N soils had different community structure than control<br />

and low N soils as predicted with OTU based and phylogenetic beta-diversity<br />

metrics. Differences in community composition and higher numbers <strong>of</strong> unknown<br />

OTUs in high N soil fungal communities may suggest that a previously uncatalogued<br />

portion <strong>of</strong> the community is released by the decline <strong>of</strong> the dominant Russulaceae.<br />

Mueller, Olaf 1 , Scott Baker 3 , Guillaume Blanc 2 , Frank Collart 3 , Fred Dietrich 1 ,<br />

Peter Larsen 3 , Jon Magnuson 3 , Francis Martin 4 , Emmanuelle Morin 4 , François<br />

Lutzoni 1 , and Daniele Armaleo 1 . 1 Department <strong>of</strong> Biology, Duke University,<br />

Durham, NC 27708 USA, 2 Information Génomique et Structurale (IGS), CNRS-<br />

UPR2589, IFR-88, Marseille, 3 Argonne National Laboratory, 9700 S. Cass Avenue,<br />

Argonne, IL 60439, 4 IUMR 1136, INRA-Nancy University, Interactions<br />

Arbres/Microorganismes, INRA-Nancy, 54280 Champenoux, France. Insights<br />

from comparative genomics in lichen symbiosis<br />

Mutual recognition and response among plants and fungi are central to either<br />

pathogenic or symbiotic associations. Genomes <strong>of</strong> the lichen-forming ascomycete<br />

Cladonia grayi (mycobiont) and its photoautotrophic symbiont, the single-celled<br />

green alga Asterochloris sp. (photobiont), were sequenced and<br />

analyzed to identify genes specific for the establishment and maintenance <strong>of</strong><br />

lichen symbiosis. Annotated gene models <strong>of</strong> C. grayi and Asterochloris were<br />

compared to reference genomes representing related Ascomycota taxa (Euro-<br />

Continued on following page


tiomycetes, Leotiomycetes, Sordariomycetes and Dothideomycetes) and Chlorophyta<br />

taxa (Chlorophyceae, Trebouxiophyceae and Prasinophyceae), respectively.<br />

A mutual best-hit blast approach was performed to identify similarities and differences<br />

in gene inventories and to provide a scaffold for detailed phylogenetic<br />

analyses. Markov cluster (MCL) studies compiled expanded and contracted gene<br />

families in the Cladonia symbionts. Gene families were investigated phylogenetically.<br />

Reduced rates <strong>of</strong> evolution driven by purifying selection are expected for<br />

genes essential to the lichen symbiosis, in order to maintain the lichen-symbiosis<br />

for a period <strong>of</strong> more than 400 million years. Candidate genes identified in this<br />

manner will be discussed in light <strong>of</strong> results from transcriptomic analyses derived<br />

from RNA isolated from the mycobiont and photobiont grown separately and together<br />

in experiments aimed at reconstituting the initial developmental interactions<br />

<strong>of</strong> the lichen symbiosis.<br />

Mujic, Alija B and Joseph W Spatafora. Department <strong>of</strong> Botany and Plant Pathology,<br />

Oregon State University, Corvallis, OR, 97331. A draft genome <strong>of</strong> Rhizopogon<br />

vesiculosus using the next generation Illumina sequencing platform<br />

The ectomycorrhizal (EM) fungi Rhizopogon vinicolor and R. vesiculosus<br />

(Boletales, Basidiomycota) are cryptic sister species <strong>of</strong> false truffles that form<br />

obligate EM relationships with Pseudotsuga menziesii (Douglas Fir). They both<br />

form unique tuberculate mycorrhizae and possess a sympatric distribution<br />

throughout their range where sampled. Previous and ongoing studies have revealed<br />

that while the sporocarp and EM morphology <strong>of</strong> these fungi may be highly<br />

similar, they possess striking life history differences. Rhizopogon vesiculosus<br />

produces larger vegetative genets and shows patterns <strong>of</strong> inbreeding within a patch<br />

size <strong>of</strong> approximately 120 meters. These trends are also reflected at the landscape<br />

scale, where R. vesiculosus has been found to display moderate population differentiation.<br />

In contrast, R. vinicolor produces smaller genets and shows no patterns<br />

<strong>of</strong> restricted gene flow within sites and only marginal population differentiation at<br />

the landscape scale. The frequency and distribution <strong>of</strong> both EM root tips and<br />

sporocarps <strong>of</strong> these species are relatively equivalent at all sites sampled and surveys<br />

<strong>of</strong> the nuclear count <strong>of</strong> basidiospores have revealed equal potential for secondary<br />

homothallism. These findings afford some explanation <strong>of</strong> the forces driving<br />

divergence between these species, but basic elements <strong>of</strong> these species’ biology<br />

such as the mating type system and biogeography remain unknown and limit our<br />

ability to test evolutionary hypotheses <strong>of</strong> sympatric and allopatric speciation. Recent<br />

advancements in high throughput sequencing technologies allow for greater<br />

applicability <strong>of</strong> genomic data for population and species-boundaries studies <strong>of</strong><br />

non-model organisms. In this study we present the partial draft genome <strong>of</strong> R.<br />

vesiculosus as sequenced on the Illumina HiSeq platform. This genome will allow<br />

for comparative analysis <strong>of</strong> gene content and mating type system with other Basidiomycota<br />

and ultimately for population/species-level genomic studies between<br />

R. vesiculosus and R. vinicolor.<br />

Nagy, Laszlo G 1 , Dimitrios Floudas 1 , Manfred Binder 1 , Igor Grigoriev 2 , Francis<br />

Martin 3 , and David S Hibbett 1 . 1 2<br />

Biology Department, Clark<br />

3<br />

University,<br />

Fungal Genomics Program Lead, DOE Joint Genome Institute, Plant Biology<br />

Department <strong>of</strong> Nancy University. Does ECM induce a clean sweep <strong>of</strong> decay related<br />

genes in Agaricomycetes?<br />

Ectomycorrhizae (ECM) are symbiotic interactions between fungi and<br />

green plants with an enormous impact on many ecosystems. The ECM association<br />

can be found in several lineages <strong>of</strong> both the Agaricomycetes and the<br />

Viridiplantae, with most <strong>of</strong> the species found in the Agaricales and Boletales in<br />

mushrooms and the Pinales the Rosidae in plants. Macroevolutionary aspects <strong>of</strong><br />

the origin <strong>of</strong> ECM formation have been studied extensively by phylogenetic comparative<br />

methods, resulting in highly discordant estimates regarding the number<br />

<strong>of</strong> gains and losses throughout the Agaricomycetes, however, its genetic basis is<br />

very poorly known. Recent and preliminary results suggest that the evolution <strong>of</strong><br />

decay-related gene families show spectacular patterns in switches between different<br />

nutritional modes <strong>of</strong> fungi. Therefore, in this study, we set out to record gene<br />

family expansions and contractions in selected ECM and saprotrophic taxa and<br />

correlate these changes with switches in nutritional mode. Gene phylogenies and<br />

comparisons <strong>of</strong> gene copy numbers reinforce predictions based on single<br />

genomes. Of the 9 gene families studied, GH6, GH7, GH61, GH12 and fPOX<br />

show considerable loss <strong>of</strong> gene copies in ECM lineages, whereas gene copy numbers<br />

<strong>of</strong> the Multicopper peroxidase, Dye decolorizing peroxidase, GH5 and GH28<br />

gene families are comparable to those in white-rot and brown rot lineages. This<br />

suggests that while some <strong>of</strong> the decay-related gene families tend to be released<br />

from selection in mycorrhizal lineages, eventually leading to extensive gene loss<br />

over longer timescales, others retain a function, presumably because these gene<br />

families confer more general functions and/or ECM lineages maintain some capability<br />

to attack wood.<br />

Neilen, Manon, Machiel E Noordeloos, Luis N Morgado, and Jozsef Geml. Nationaal<br />

Herbarium Nederland, Nederlands Centrum voor Biodiversiteit Naturalis,<br />

Universiteit Leiden, PO Box 9514, 2300RA Leiden, The Netherlands. Willows -<br />

the arctic connection: biodiversity <strong>of</strong> arctic-alpine ectomycorrhizal fungi in<br />

Salix repens communities in Dutch North Sea sand dunes<br />

Dwarf willow (Salix) species are ubiquitous in the Arctic and are the primary<br />

hosts for the majority <strong>of</strong> arctic ectomycorrhizal (ECM) fungi. In Western<br />

Europe, the sand dune communities <strong>of</strong> creeping willow (S. repens) are known to<br />

harbour a rich myc<strong>of</strong>lora and are highly important for nature conservation, water<br />

resource management, and recreational purposes. In this project, we assess the<br />

biodiversity <strong>of</strong> ECM fungi in the Arctic (Alaska and Svalbard) and in Dutch<br />

coastal dune communities. We generated DNA sequence data from both soil and<br />

fruitbody samples and conducted phylogenetic analyses for the major Salix-symbiont<br />

ECM genera, such as Cortinarius, Russula, Hebeloma, Inocybe and Tomentella.<br />

Our results show that the majority <strong>of</strong> the ECM basidiomycete species<br />

found in soil samples taken from S. repens beds in the Dutch maritime sand dunes<br />

also occur in various arctic tundra communities. When comparing our results with<br />

long-term records <strong>of</strong> aboveground diversity (i.e., sporocarp mapping data generated<br />

by the Werkgroep Paddenstoelenkartering Nederland, WPN), we found that<br />

many <strong>of</strong> these ‘arctic’ species are considered rare and/or threatened, and therefore,<br />

red-listed in the Netherlands. In addition, we could confidently identify several<br />

OTUs to arctic species that had never been reported from the Netherlands before,<br />

e.g., Cortinarius favrei, Cortinarius laetissimus, Inocybe straminipes, Russula<br />

nana. Based on the many similarities in the ECM communities in arctic tundra<br />

and maritime sand dunes, it appears that, in the temperate zone, S. repens may<br />

serve as an ecological analog <strong>of</strong> arctic dwarf willows. In coastal dunes, S. repens<br />

is among the very few ECM hosts and its presence strongly enhances fungal biodiversity<br />

in these areas. Knowledge on ECM biodiversity can provide insight into<br />

the functioning and sustainability <strong>of</strong> these ecosystems.<br />

Nguyen, Nhu H 1 , Karen Hansen 2 , Rosanne Healy 3 , and Else C Vellinga 1 . 1<br />

Dept. <strong>of</strong> Plant and Microbial Biology, 111 Koshland Hall, UC Berkeley, Berkeley<br />

CA 94720, USA, 2 Swedish Museum <strong>of</strong> Natural History, Department <strong>of</strong> Cryptogamic<br />

Botany, PO Box 50007, SE-104 05 Stockholm, Sweden, 3 Department <strong>of</strong><br />

Plant Biology, University <strong>of</strong> Minnesota, St. Paul, MN 55108, USA. Helvella vespertina<br />

and Helvella dryophila, two new Helvella lacunosa look-alikes from<br />

western North <strong>America</strong><br />

Fungal taxonomists have long suspected that the macr<strong>of</strong>ungi <strong>of</strong> western<br />

North <strong>America</strong> are distinct from eastern North <strong>America</strong> and Europe. Yet, we continue<br />

to apply European species name to western North <strong>America</strong>n species. One<br />

such case is Helvella “lacunosa”, a very common ectomycorrhizal species. We<br />

collected and sequenced ITS and LSU rRNA genes <strong>of</strong> Helvella “lacunosa” from<br />

California and compared them to European and eastern North <strong>America</strong>n H. lacunosa.<br />

Specimens from western North <strong>America</strong> are distinct from European specimens<br />

and from <strong>America</strong>n specimens east <strong>of</strong> the Rocky Mountains. Furthermore,<br />

the western North <strong>America</strong>n “lacunosa” resolved into two species, one occurring<br />

with conifers (pines and Douglas fir) and one occurring with hardwood, mostly<br />

oak. These two species will be described as Helvella vespertina and Helvella<br />

dryophila. Intensive sampling in eastern North <strong>America</strong> and Mexico is desperately<br />

needed to resolve the H. “lacunosa” complex occurring in North <strong>America</strong>.<br />

This effort is a contribution to the North <strong>America</strong>n Myc<strong>of</strong>lora Project.<br />

Nguyen, Hai, Nancy Nickerson, and Keith Seifert. Agriculture and Agri-Food<br />

Canada | 960 Carling Avenue | Ottawa, Ontario, Canada | K1A 0C6. Basidioascus:<br />

a new lineage <strong>of</strong> heat resistant and xerotolerant basidiomycetes<br />

Basidioascus was originally described from soil in Australia by Matsushima,<br />

who considered the type species, B. undulatus, to be an ascomycete with<br />

some basidiomycete characters. We isolated this fungus commonly from soil in<br />

temperate areas or Canada and the United States, using methods designed to isolate<br />

heat resistant and xerotolerant fungi. Molecular phylogenetic analysis <strong>of</strong><br />

rDNA loci reveals that Basidioascus is a member <strong>of</strong> the Basidiomycota, with a<br />

sister-group relationships with Wallemia, a xerotolerant basidiomycetous mould<br />

that contaminates food and house dust. Microscopy and molecular phylogenetic<br />

analysis <strong>of</strong> the internal transcribed spacer (ITS) barcode region support the recognition<br />

<strong>of</strong> B. undulatus and two undescribed species. Our interpretation <strong>of</strong> microscopic<br />

observations is that ontogenesis starts with the development <strong>of</strong> lateral or<br />

terminal clavate, probasidia on hyphae, which become obovate when mature.<br />

After self-fertilization, the basidium ruptures, producing 1-3 dark, thick-walled<br />

globose to subglobose basidispores near the surface <strong>of</strong> the agar, leaving behind a<br />

collapsed basidium.<br />

Njambere, Evans N 1 , Frank Wong 2 , and Ning Zhang 1 . 1 Dept. <strong>of</strong> Plant Biology<br />

and Pathology, Rutgers University, New Brunswick, NJ 08901, 2 Bayer Environmental<br />

Science, Alexandria, VA. Microsatellite markers reveal isolation by<br />

distance and recent population expansion <strong>of</strong> Waitea circinata var. circinata<br />

Waitea circinata var. circinata (Wcc) is an emerging pathogen <strong>of</strong> turf<br />

grass in North <strong>America</strong>. It causes brown ring patch <strong>of</strong> turf grass in golf courses<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 35


and amenity areas. The population dynamics and dispersal pathway <strong>of</strong> this<br />

pathogen in the US have not been well studied. In attempt to elucidate this, we<br />

isolated eight promising microsatellite markers from an enriched genomic library<br />

<strong>of</strong> Wcc. Seven <strong>of</strong> these markers were used to study the population diversity <strong>of</strong><br />

eastern and western US populations <strong>of</strong> Wcc and the association with phenotypic<br />

characteristics. In general, the microsatellite markers were highly polymorphic,<br />

with an observed heterozygosity <strong>of</strong> > 0.5. The eastern population was differentiated<br />

from the western population but both populations showed the presence <strong>of</strong> immigrant<br />

individuals across boundaries. Analysis <strong>of</strong> population structure with geography<br />

revealed a significant clinal expansion, r 2 = 0.11, P = 0.01, <strong>of</strong> Wcc within<br />

a radius <strong>of</strong> about 150 km. However, it became more stochastic with increase in<br />

distance. No host specific clusters were observed based on UPGMA clustering,<br />

suggesting a lack <strong>of</strong> association between host and pathogen genotypes.<br />

Norvell, Lorelei L 1 and Scott A Redhead 2 . 1 Mycotaxon, PNW Mycology<br />

Service, Portland OR 97229-1309 USA, 2 National <strong>Mycological</strong> Herbarium, Eastern<br />

Cereal & Oilseed Research Centre C.E.F., Agriculture & Agri-Food Canada,<br />

Ottawa, ON K1A 0C6 Canada. A mycologist’s guide to the Melbourne Code.<br />

The International Code <strong>of</strong> Nomenclature for algae, fungi, and plants<br />

(ICN) replaced the International Code <strong>of</strong> Botanical Nomenclature (ICBN) during<br />

the 2011 International Botanical Congress (IBC) in Melbourne. The new ICN permits<br />

diagnoses/descriptions in English (no longer requiring Latin) and electronic<br />

publication <strong>of</strong> fungal names (via PDFs in ISSN/ISBN numbered publications).<br />

With registration <strong>of</strong> fungal names required for valid publication in 2013, the IBC<br />

Nomenclature Committee for Fungi (NCF) is now moving to approve one or<br />

more nomenclatural registries. The ICN also excludes names <strong>of</strong> microsporidians<br />

and organisms treated in other Codes, refines typification procedures and spelling<br />

<strong>of</strong> sanctioned names, and recommends how to designate type cultures. The Melbourne<br />

Congress also referred mechanical typification issues and governance <strong>of</strong><br />

fungal nomenclature to special committees and approved NCF recommendations<br />

on the conservation and rejection <strong>of</strong> notable fungal names. The new ICN does not<br />

permit multiple names for fungi with pleomorphic life cycles, which the ICBN<br />

(under Art. 59) previously allowed for non-lichenized ascomycetes and basidiomycetes.<br />

After 2013, alternative new names will be either illegitimate or not<br />

valid, although alternatives published through 2012 may be valid or legitimate.<br />

Although teleomorph- and anamorph-typified names will compete equally for priority,<br />

the ICN dictates delay in adopting anamorph-typified names with priority<br />

over commonly used teleomorph-typified names. To buffer the effects <strong>of</strong> this<br />

major change regarding alternative names, NCF-sanctioned subcommittees will<br />

compile and adjudicate lists <strong>of</strong> conserved/rejected names <strong>of</strong> non-lichenized fungi<br />

(not restricted to the Ascomycota and Basidiomycota) for approval by the IBC.<br />

The NCF and International Commission on the Taxonomy <strong>of</strong> Fungi (ICTF) - collaborating<br />

to issue joint guidelines, updates, and advice - recognize that mycological<br />

opinion over the Art. 59 change is still deeply divided and must be considered.<br />

Nuhn, Mitchell, Roy E Halling, Manfred Binder, David S Hibbett, and Todd W<br />

Osmundson. Department <strong>of</strong> Biology, Clark University, 950 Main St, Worcester,<br />

MA 01610. The Boletineae <strong>of</strong> Queensland, Australia<br />

The Boletineae is a group <strong>of</strong> mushrooms with many species highly sought<br />

after by both animals and humans as food. The majority <strong>of</strong> Boletineae species are<br />

ectomycorrhizal with woody plants, making them important members <strong>of</strong> forest<br />

ecosystems. This study will provide the first biodiversity inventory <strong>of</strong> the Boletineae<br />

in Queensland, Australia, including three <strong>of</strong> Australia’s Biodiversity<br />

Hotspots and World Heritage listed sites. To accomplish the inventory, modern<br />

molecular and phylogenetic techniques will be used, in combination with traditional<br />

taxonomic methods. A sequence dataset <strong>of</strong> nuclear ribosomal large subunit<br />

(nuclsu) and translation elongation factor 1-alpha (tef1) sequences has been generated<br />

for the systematic overview <strong>of</strong> the Boletineae <strong>of</strong> Queensland. The study has<br />

already resulted in one publication, designating a new cosmopolitan genus and at<br />

least two species for a previously described species, Sutorius eximius and S. australianses<br />

(previously Tylopilus eximius). Further study is ongoing, providing an<br />

overview <strong>of</strong> amorphologically distinct, cosmopolitan taxon, Tylopilus chromapes,<br />

that appears to be polyphyletic based on preliminary results. An overview <strong>of</strong> additional<br />

Boletineae species occurring in Queensland, Australia and their relationships<br />

in the context <strong>of</strong> a worldwide phylogeny <strong>of</strong> Boletineae is in progress.<br />

Olarte, Rodrigo A 1 , Bruce W Horn 2 , Carolyn J Worthington 1 , and Ignazio Carbone<br />

1 . 1 Department <strong>of</strong> Plant Pathology, North Carolina State University,<br />

Raleigh, NC 27695, 2 National Peanut Research Laboratory, Agricultural Research<br />

Service, U.S. Department <strong>of</strong> Agriculture, Dawson, GA 39842. Matingtype<br />

heterokaryosis and population shifts in Aspergillus flavus<br />

Aspergillus flavus is a fungal pathogen <strong>of</strong> many agronomically important<br />

crops worldwide. We sampled A. flavus strains from a cornfield in Rocky Mount,<br />

NC. This field was planted in 2010 and plots were inoculated at tasseling with either<br />

AF36 or NRRL 21882 (=Afla-Guard) biocontrol strains, both <strong>of</strong> which are<br />

36 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

mating type MAT1-2. Subsequently, toxigenic strain NRRL 3357 (MAT1-1) was<br />

applied to all plots, including control plots not inoculated with biocontrol strains.<br />

Sclerotia were harvested from infected corn ears approximately 4.5 months after<br />

planting (2.5 months after biocontrol treatment); ninety single-ascospore isolates<br />

were isolated from ascocarps originating from plots treated with AF36 and NRRL<br />

21882. In addition, eighty A. flavus isolates were collected from soil one month<br />

after planting (before biocontrol application) and one year after biocontrol application,<br />

for a grand-total <strong>of</strong> 250 isolates. Aflatoxin (AF) and cyclopiazonic acid<br />

(CPA) production were determined using standard thin-layer chromatography<br />

and HPLC. Three distinct toxin classes were identified: AF-/CPA-, AF+/CPA+<br />

and AF-/CPA+. PCR amplification revealed grouping <strong>of</strong> isolates into three distinct<br />

mating-type classes: MAT1-1, MAT1-2 and MAT1-1/MAT1-2. A significant<br />

proportion (54%) <strong>of</strong> isolates sampled prior to biocontrol treatments were heterokaryotic<br />

for mating type (MAT1-1/MAT1-2), and 39% <strong>of</strong> isolates obtained<br />

from ascospores were heterokaryotic as well as 9% <strong>of</strong> isolates from soil after biocontrol<br />

treatments. The vertical transmission <strong>of</strong> MAT1-1/MAT1-2 to progeny ascospore<br />

isolates suggests that heterokaryosis can be maintained in subsequent<br />

generations. The population genetic structure before and after the application <strong>of</strong><br />

biocontrol treatments will be discussed. Further characterization <strong>of</strong> heterokaryons<br />

and their frequency in A. flavus populations may be important in understanding<br />

the adaptation <strong>of</strong> these fungi to changing environmental conditions.<br />

Oliver, Jason P and Jonathan S Schilling. University <strong>of</strong> Minnesota, 320 Kaufert<br />

Laboratory, 2004 Folwell Avenue, St. Paul, MN 55108-6130. Manure gas eating<br />

microbes: Effect <strong>of</strong> fungal and bacterial biomass on bi<strong>of</strong>iltration <strong>of</strong> livestock<br />

production emissions<br />

Though specific data is only just being analyzed, scientists and regulators<br />

are beginning to recognize the significant emission <strong>of</strong> particulate matter, odor,<br />

hazardous and greenhouse gases from livestock production and manure management.<br />

Bi<strong>of</strong>iltration - use <strong>of</strong> a biologically active porous media to capture and degrade<br />

gaseous pollutants - is a low-cost and adaptable mitigation technology suitable<br />

for manure storage emissions. Improving the design and management <strong>of</strong><br />

bi<strong>of</strong>ilters (and ultimately adoption by producers) rests on our ability to understand<br />

the microbial ecology <strong>of</strong> the filter media, the underpinnings <strong>of</strong> bi<strong>of</strong>ilter effectiveness<br />

and efficiency. Studying bi<strong>of</strong>ilters at multiple scales, we are beginning to resolve<br />

the effect <strong>of</strong> fungal and bacterial biomass on reductions <strong>of</strong> gases generated<br />

by long-term storage <strong>of</strong> livestock manure.<br />

Orwin, Kate H 1 , Miko U F Kirschbaum 2 , Julie R Deslippe 2 , and Ian A Dickie<br />

3 . 1 Lancaster Environment Centre, Lancaster University, Lancaster, LA1 4YQ,<br />

United Kingdom, 2 Landcare Research, Private Bag 11052, Palmerston North<br />

4442, New Zealand, 3 Landcare Research, Box 40, Lincoln 7640, New Zealand.<br />

The impact <strong>of</strong> mycorrhizal fungal traits on ecosystem responses to global<br />

change: a modeling approach<br />

Ecosystems are currently being subjected to multiple global change factors,<br />

from nitrogen deposition to increasing CO2 concentrations and changes in<br />

temperature. Understanding the consequences <strong>of</strong> these effects on ecosystem<br />

processes such as carbon cycling is <strong>of</strong> prime importance, as this will determine<br />

whether ecosystems are likely to cause positive or negative feedbacks to climate<br />

change. Current evidence suggests that carbon cycling can be strongly affected by<br />

the traits <strong>of</strong> the plant community, with most studies focusing on litter quality and<br />

plant growth rates. However, recent modelling work has demonstrated that the<br />

traits <strong>of</strong> plant symbionts can also play a significant role in determining ecosystem<br />

carbon storage. In particular, the ability <strong>of</strong> mycorrhizal fungi to directly access organic<br />

nutrients can be as important as a major shift in litter quality for carbon storage.<br />

The two main mechanisms behind this effect were direct access to organic<br />

nutrients causing i) an increase in plant nutrition resulting in enhanced photosynthetic<br />

rates and subsequent carbon inputs to the soil and ii) a reduction in microbial<br />

access to nutrients, which reduced the rate <strong>of</strong> soil organic matter decomposition.<br />

We have little understanding <strong>of</strong> the way in which such interactions among<br />

mycorrhizal symbionts, their hosts and other soil decomposers may modify the effect<br />

<strong>of</strong> global change factors on ecosystem carbon dynamics. Here, we use a<br />

newly developed model (MySCaN: Mycorrhizal Status, Carbon and Nutrient cycling)<br />

to examine the impact <strong>of</strong> global change factors on carbon storage, and how<br />

these trends can be affected by mycorrhizal traits. We show that global change effects<br />

on carbon storage are not universal, but are modified by the interaction between<br />

plants and their mycorrhizal symbionts.<br />

Osmundson, Todd W and Matteo Garbelotto. Forest Mycology and Pathology<br />

Laboratory, Department <strong>of</strong> Environmental Science, Policy and Management,<br />

University <strong>of</strong> California, Berkeley, CA 94720, USA. Fire drives habitat specialization,<br />

speciation, and hybridization <strong>of</strong> Morel mushrooms in western<br />

North <strong>America</strong><br />

Continued on following page


Morels (Morchella spp.) are commercially valuable edible mushrooms<br />

and non-timber forest products. Several species reproduce prolifically the year<br />

after wildfires, suggesting a response to - and perhaps important role in - post-fire<br />

forest regeneration. Although post-fire seed regeneration may accelerate evolutionary<br />

diversification compared to adult persistence in plants, similar comparisons<br />

for fungi are lacking. Using multilocus phylogenies and diversification<br />

analyses, we demonstrate the evolutionary importance <strong>of</strong> fire-associated reproduction<br />

in morels through (i) high ecological specificity <strong>of</strong> fire association; (ii)<br />

phylogenetic conservatism <strong>of</strong> fire-association, with diversification concurrent<br />

with shifts in fire-association status; and (iii) gene introgression between phylogenetically<br />

distinct but ecologically similar fire-associated species. Rates <strong>of</strong> phylogenetic<br />

diversification and molecular evolution do not differ between fire- and<br />

non-fire-associated lineages, suggesting that rate-enhancing factors (e.g., population<br />

turnover, large population sizes) are counteracted by rate-limiting ones (e.g.,<br />

long fire-return intervals). Genetic introgression was observed only between fireassociates,<br />

suggesting that fire can counteract diversification by synchronizing<br />

sexual reproduction between species, a process potentially relevant to other<br />

species that reproduce in response to disturbance.<br />

Owensby, Alisha C, Kathryn E Bushley, and Joseph W Spatafora. Department<br />

<strong>of</strong> Botany and Plant Pathology, Oregon State University, Corvallis, OR. Comparative<br />

genomics reveals complex patterns <strong>of</strong> secondary metabolite evolution<br />

in the genus Elaphocordyceps<br />

Hypocreales is an order characterized by a dynamic evolutionary history<br />

<strong>of</strong> interkingdom host jumping, with members that parasitize animals, plants, and<br />

other fungi. The monophyly <strong>of</strong> taxa attacking members <strong>of</strong> the same kingdom has<br />

not been supported by molecular phylogenetics, however. For example, Trichoderma<br />

spp. and Elaphocordyceps spp. are both mycoparasitic, but are members<br />

<strong>of</strong> two different families within the Hypocreales, the Hypocreaceae and Ophiocordycipitaceae,<br />

respectively. In fact, both species are more closely related to insect<br />

pathogens, than they are to each other. Three species <strong>of</strong> Trichoderma have<br />

sequenced genomes, and more recently the genomes <strong>of</strong> several insect pathogens<br />

in the Hypocreales have been published (e.g. Cordyceps militaris, Metarhizium<br />

anisopliae, M. acridum) and others are in progress (e.g. Tolypocladium inflatum).<br />

Elaphocordyceps is a genus within this clade that primarily parasitizes ectomycorrhizal<br />

truffles in the genus Elaphomyces, representing a yet unsampled ecology<br />

within Hypocreales. To compare gene space <strong>of</strong> a truffle pathogen with closely related<br />

insect pathogens and distantly related mycoparasites, we sequenced the<br />

genome <strong>of</strong> Elaphocordyceps ophioglossoides. Our draft assembly <strong>of</strong> the E.<br />

ophioglossoides genome has a size <strong>of</strong> approximately 32 MB and 10,779 predicted<br />

genes. Here, we present a survey <strong>of</strong> E. ophioglossoides genes involved in secondary<br />

metabolism. Notably, E. ophioglossoides is missing the cyclosporin producing<br />

nonribosomal peptide synthetase (NRPS) found in its congener, T.<br />

inflatum, but possesses the destruxin NRPS, recently identified in Metarhizium<br />

anisopliae. The three largest genes in E. ophioglossoides are related to peptaibol<br />

NRPS genes, which are only described from Trichoderma spp., and are indicated<br />

as being important to mycoparasitism. Our work represents the first genome <strong>of</strong> an<br />

Elaphocordyceps species parasitic on other fungi and will provide insights into<br />

mechanisms underlying mycoparasitism and interkingdom host jumping.<br />

Padamsee, Mahajabeen 1 , Stanley E Bellgard 2 , Ian Dickie 3 , Scott Fraser 4 ,<br />

Duckchul Park 1 , Bryan Stevenson 4 , and Bevan Weir 1 . 1 Biosystematics Team,<br />

Landcare Research, Auckland, New Zealand, 2 Biodiversity and Conservation<br />

Team, Landcare Research, Auckland, New Zealand, 3 Ecosystem Processes,<br />

Landcare Research, Lincoln, New Zealand, 4 Soils and Landscape Responses,<br />

Landcare Research, Hamilton, New Zealand. Stuff in Duff: a metagenomic<br />

study <strong>of</strong> fungi in kauri forests<br />

Kauri (Agathis australis, Araucariaceae) is restricted in distribution to the<br />

northern tip <strong>of</strong> the North Island <strong>of</strong> New Zealand. Living to 2,500 years or more<br />

and having trunks to 7 m diam., kauri is a keystone species. Kauri was extensively<br />

logged and now less than 1% remains <strong>of</strong> undisturbed old-growth forests. Since<br />

the 1970s these trees have also been under threat from the exotic invasive, Phytophthora<br />

taxon Agathis (PTA) that causes kauri dieback. We are interested in understanding<br />

how the fungal community associated with the large amount <strong>of</strong> kauri<br />

duff (i.e., leaf litter) will be affected after infection <strong>of</strong> the tree by this exotic root<br />

pathogen. However, there has been no thorough study <strong>of</strong> the fungi associated with<br />

kauri and as a result we have no baseline data <strong>of</strong> the composition <strong>of</strong> a “healthy”<br />

fungal community in kauri leaf litter. We used pyrosequencing to study fungal<br />

biodiversity and assess changes in fungal community composition after invasion<br />

by PTA. The sampling scheme was optimized and several methods <strong>of</strong> analyzing<br />

the data were tested. A sequence database <strong>of</strong> known fungi was created with which<br />

to compare the metagenomic data.<br />

Padamsee, Mahajabeen and Eric McKenzie. Biosystematics Team, Landcare<br />

Research, Auckland, New Zealand. Are there New Zealand-only clades <strong>of</strong> rust<br />

fungi?<br />

Of the 250 species <strong>of</strong> rust fungi (Pucciniales) recorded in New Zealand,<br />

approximately 90 are believed to be endemic since they occur on native plants.<br />

Over the last 150 years, the number <strong>of</strong> non-native rusts has risen sharply from 33<br />

to over a hundred. Although most <strong>of</strong> these rusts are recorded on non-native plants,<br />

several have been also recorded on native hosts. This begs the question, if host<br />

jumping has been observed in such a short time, what implication does this have<br />

for the origin <strong>of</strong> endemic rusts, i.e., are they really endemic? If not, then when and<br />

from where were the rusts introduced? Recent evidence suggests that many <strong>of</strong> the<br />

non-native rusts have been introduced from Australia by trans-Tasman airflows,<br />

which suggests that historically dispersal from Australia may have influenced the<br />

current distribution <strong>of</strong> rust fungi. A multi-gene phylogeny <strong>of</strong> New Zealand rust<br />

fungi was constructed and when combined with data from non-native rusts suggested<br />

the strong likelihood that there are New Zealand-only clades <strong>of</strong> rust fungi.<br />

Preliminary results indicate that several rust fungi may have evolved on closely<br />

related hosts in New Zealand.<br />

Pagliaccia, Deborah P, Elinor P Pong, Brandon M McKee, and Greg W Douhan.<br />

Department <strong>of</strong> Plant Pathology and Microbiology, University <strong>of</strong> California,<br />

Riverside, CA 92521. Population genetic structure <strong>of</strong> Phytophthora cinnamomi<br />

Rands associated with Phytophthora root rot <strong>of</strong> avocado within California<br />

and the potential introduction <strong>of</strong> a new clonal lineage<br />

Phytophthora root rot (PRR) <strong>of</strong> avocado (Persea americana Mill), caused<br />

by Phytophthora cinnamomi, is the most serious disease <strong>of</strong> avocado worldwide.<br />

Previous studies have determined that this pathogen exhibits a primarily clonal reproductive<br />

mode but no population level studies have been conducted in the avocado<br />

growing regions <strong>of</strong> California (Ca). Therefore, we used AFLP and matingtype<br />

to investigate pathogen diversity from 138 isolates collected in 2009-2010<br />

from 15 groves from the Northern and Southern avocado growing regions. Additional<br />

isolates collected from avocado from 1970 to 2007 as well as isolates from<br />

other countries and hosts were also used for comparative purposes. Three distinct<br />

clades were found based on UPGMA analysis <strong>of</strong> 22 polymorphic loci. One clade<br />

contained only A1 isolates collected from various hosts and locations. Two A2<br />

clades were found; one clade contained both newer and older collections from<br />

Northern and Southern Ca whereas the other clade only contained isolates collected<br />

in 2009-10. A total <strong>of</strong> 16 genotypes were found with only 1 to 4 genotypes<br />

identified from any one location. The results indicate significant population structure<br />

in the Ca avocado P. cinnamomi population, low genotypic diversity consistent<br />

with asexual reproduction, potential evidence for the movement <strong>of</strong> clonal<br />

genotypes between the two growing regions, and a potential introduction <strong>of</strong> a new<br />

clonal lineage into Southern CA. Current studies are underway to investigate the<br />

phenotypic variation (in vitro and in vivo) within our collection <strong>of</strong> P. cinnamomi<br />

isolates to determine if these new findings will affect our efforts towards breeding<br />

PRR tolerant rootstocks <strong>of</strong> avocado to control this important disease.<br />

Perry, Brian A, Erin Datl<strong>of</strong>, and Mali‘o Kodis. Biology Department, University<br />

<strong>of</strong> Hawai‘i at Hilo, 200 W. Kawili St., Hilo, HI 96720. Foliar endophytic community<br />

structure in wild and cultivated stands <strong>of</strong> Hawaiian ‘ohi’a lehua<br />

(Metrosideros polymorpha)<br />

Although associated with all plants that have been investigated, the taxonomic,<br />

genetic and functional diversity <strong>of</strong> endophytic fungi remains undocumented<br />

for many regions, including the Hawaiian Islands. Fungal endophytes<br />

have been shown to confer such benefits as increased draught tolerance, resistance<br />

to pathogens, and anti-herbivory properties. Additionally, fungal endophyte composition<br />

has been shown to play a large role in altering vegetation dynamics and<br />

plant community composition. Given the potential roles <strong>of</strong> fungal endophytes, it<br />

is clear that a detailed understanding the taxonomic and genetic diversity <strong>of</strong> these<br />

symbionts should be addressed as conservation and management plans for host<br />

species are developed. To assess such diversity in a native Hawaiian taxon, we examined<br />

foliar fungal endophytic community structure in elevational phenotypes<br />

<strong>of</strong> wild and cultivated Metrosideros polymorpha (‘ohi’a lehua) using environmental<br />

PCR methods. Preliminary results based on ITS sequence data indicate<br />

that low and high elevation populations on Hawai’i Island harbor disparate endophytic<br />

communities. Additionally, seeds collected from high and low elevation<br />

populations and reared together in a mid-elevation garden harbor endophytic<br />

communities similar to those <strong>of</strong> their parent populations, suggesting strong host<br />

selection and/or vertical symbiont transmission. Seeds isolated from cultivated M.<br />

polymorpha individuals are currently being examined as potential vectors <strong>of</strong> endophytic<br />

communities similar to those recovered from foliar samples.<br />

Petersen, Ronald H and Karen W Hughes. Ecology & Evolutionary Biology,<br />

University <strong>of</strong> Tennessee, Knoxville, Tn 37996-1100. The “Big Ditch” project:<br />

TransAtlantic mushroom disjunction tested using multiple taxonomic tools<br />

Written descriptions and illustrations by historical Euro-Scandinavian<br />

mycologists crossed the Atlantic to <strong>America</strong>, where they were matched to fungi<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 37


<strong>of</strong> the New World. Doubts about the accuracy <strong>of</strong> name applications across the<br />

ocean have persisted, but molecular phylogenies now provide a new level <strong>of</strong> resolution.<br />

Our research targets fleshy fungi whose names originated in Europe, but<br />

which bear the same name and superficial form in eastern North <strong>America</strong>. Several<br />

examples can be identified in which DNA sequences reveal differences across<br />

the ocean, including saprophytic (Marasmius rotula; Baeospora myosura;<br />

Sparassis crispa) and ectomycorrhizal (Strobilomyces strobilaceus; Tricholoma<br />

populinum) fungi. In addition, it appears that genes for sexual compatibility and<br />

recognition are more highly conserved than the internally transcribed spacer<br />

(ITS), but morphological differences vary with the individual taxon. Implications<br />

<strong>of</strong> this research include the necessity for new names for heret<strong>of</strong>ore cryptic taxa on<br />

both continents and resultant adjustment <strong>of</strong> biodiversity inventory reports.<br />

Peterson, Stephen W. USDA-ARS NCAUR, 1815 North University Street, Peoria,<br />

IL 61604. Aspergillus section Versicolores: nine new species and multilocus<br />

DNA sequence based phylogeny<br />

ß-tubulin, calmodulin, internal transcribed spacer and partial lsu-rDNA,<br />

RNA polymerase, DNA replication licensing factor Mcm7, and pre-rRNA processing<br />

protein Tsr1 were amplified and sequenced from 62 A. versicolor clade<br />

isolates and analyzed phylogenetically using the concordance model to establish<br />

species boundaries. We used phylogeny to define species and phenotype from 10d<br />

CYA cultures to describe the species. Aspergillus austroafricanus, A. creber, A.<br />

cvjetkovicii, A. fructus, A. jensenii, A. puulaauensis, A. subversicolor, A. tennesseensis<br />

and A. venenatus are described as new species and A. amoenus, A. protuberus,<br />

A. tabacinus and A. versicolor are accepted as distinct species on the<br />

basis <strong>of</strong> molecular and phenotypic differences. PCR primer pairs used to detect A.<br />

versicolor in sick building syndrome studies have a positive reaction for all <strong>of</strong> the<br />

newly described species.<br />

Phookamsak, Rungtiwa 1 , Amy Y Rossman 2 , Ekachai Chukeatirote 1 , and<br />

Kevin D Hyde 1 . 1 School <strong>of</strong> science, Mae Fah Luang University, 57100, Chiang<br />

Rai, Thailand., 2 Mycology & Microbiology Laboratory, USDA-ARS, Beltsville,<br />

MD 20705, USA. Taxonomic and phylogenetic studies <strong>of</strong> Dothideomycetes<br />

on bamboo in Chiang Rai Province, Northern <strong>of</strong> Thailand<br />

Bamboos are fascinating and useful plants that have a wide range <strong>of</strong> uses.<br />

The traditional people in Southeast Asia, China, Japan, India and South <strong>America</strong><br />

have used bamboos in their cultures and survival since ancient times. There are<br />

several studies on endophytic, pathogenic and saprobic fungi on bamboo. However,<br />

phylogenetic information based on molecular data is poorly known for many<br />

bambusicolous fungi. We studied the taxonomy and phylogeny <strong>of</strong> Dothideomycetes<br />

on bamboo in Chiang Rai Province, Northern Thailand. Specimens<br />

were collected from six locations in Chiang Rai. Pure cultures were obtained<br />

using single spore isolation. We identified bitunicate fungi on dead stems or<br />

branches <strong>of</strong> bamboo that belong to the families <strong>of</strong> Aigialaceae, Didymosphaeriaceae,<br />

Lophiostomataceae, Massarinaceae and Melanommataceae. Some fungi<br />

are illustrated and a checklist <strong>of</strong> fungi examined on bamboo is presented. Among<br />

the 33 saprobic fungi on dead materials two isolates <strong>of</strong> Aigialaceae, four <strong>of</strong> Didymosphaeriaceae,<br />

three <strong>of</strong> Lophiostomataceae and Massarinaceae and 11 <strong>of</strong><br />

Melanommataceae were identified. The molecular phylogeny <strong>of</strong> Dothideomycetes<br />

on bamboo is inferred based on combined sequences <strong>of</strong> rDNA, the<br />

internal transcribed spacers (ITS rDNA), small subunit nuclear rDNA (18S,<br />

SSU), large subunit nuclear rDNA (28S, LSU) and partial RNA polymerase second<br />

largest subunit (RPB2).<br />

Picard, Kathryn T 1 , Rowena F Stern 2 , and François Lutzoni 1 . 1 Duke University,<br />

Durham, NC 27708, 2 Sir Alister Hardy Foundation for Ocean Science, Plymouth,<br />

UK. Investigating early-diverging fungi from marine and estuarine<br />

habitats in North <strong>America</strong> and Europe<br />

Despite increasing efforts to characterize and catalog early-diverging<br />

fungi, marine and estuarine habitats remain largely unexplored. Further, traditional<br />

culturing methods commonly employed in taxonomic surveys provide only<br />

limited insight into the breadth <strong>of</strong> zoosporic fungal diversity. However, with the<br />

application <strong>of</strong> culture-independent molecular techniques, including environmental<br />

cloning and, increasingly, high-throughput sequencing, the presence, distribution,<br />

and variety <strong>of</strong> these previously unculturable microbes is being elucidated.<br />

For this study, marine and estuarine habitats in North <strong>America</strong> and Europe were<br />

surveyed for novel fungal phylotypes using next-generation sequencing methods.<br />

North <strong>America</strong>n samples originated from a yearlong sampling effort in coastal<br />

North Carolina. Monthly sediment and plankton samples were collected from<br />

three small estuarine islands located within the Beaufort Inlet in Pamlico Sound.<br />

Quarter-annual collections <strong>of</strong> marine sediments from the shallow waters <strong>of</strong> Cape<br />

Lookout Bight were also made. Using the Ion Torrent platform, the 5’ end <strong>of</strong> the<br />

nuclear large subunit (28S) was sequenced from these samples, resulting in ~200<br />

bp reads. European samples originated from open water collections made across<br />

the English Channel over the course <strong>of</strong> three months. Partial nuclear small subunit<br />

(18S) rDNA sequences (~200-400 bp) were generated from English Channel<br />

38 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

samples using 454 pyrosequencing. The conserved 18S and 28S markers were<br />

chosen over more variable regions, such as ITS1 and ITS2, based on greater availability<br />

<strong>of</strong> publicly accessible reference sequences. Preliminary analysis <strong>of</strong> both<br />

North <strong>America</strong>n and European sequence data reveals novel fungal phylotypes<br />

from across the fungal tree.<br />

Porras-Alfaro, Andrea 1,4 , Liu Kuan-Liang 3 , Zachary Gossage 1 , Lynnaun<br />

Johnson 1 , Tabitha Williams 1 , Gary Xie 2 , and Cheryl R Kuske 2 . 1 Department <strong>of</strong><br />

Biological Sciences, Western Illinois University, Illinois, USA, 2 Los Alamos National<br />

Laboratory, Bioscience Division, Los Alamos, New Mexico, USA, 3 Institute<br />

<strong>of</strong> Information Management, National Cheng Kung University, Tainan City,<br />

Taiwan, Republic <strong>of</strong> China, 4 Department <strong>of</strong> Biology, University <strong>of</strong> New Mexico,<br />

Albuquerque, USA. ITS and LSU automated classification: the fungal RDP<br />

naïve Bayesian classifier<br />

The introduction <strong>of</strong> next generation sequencing (NGS) in ecological studies<br />

has created a major revolution in fungal ecology. Analyses <strong>of</strong> large fungal<br />

datasets are common but currently limited by the lack <strong>of</strong> databases and reliable<br />

tools that will allow the classification <strong>of</strong> thousands <strong>of</strong> fungal sequences from environmental<br />

samples. The main objective <strong>of</strong> this project was to create curated<br />

databases for the Internal Transcribed Spacer (ITS) region and the large-subunit<br />

rRNA (LSU) gene and to evaluate their performance in the Ribosomal Database<br />

Project (RDP) naïve Bayesian classifier. We created and compared hand-curated<br />

LSU and ITS databases (more than 8 000 fungal sequences per database). Taxonomic<br />

gaps in fungal databases and challenges in the creation and maintenance <strong>of</strong><br />

these databases will be discussed. When compared with the traditional BLASTN<br />

approach, the RDP fungal classifier was more rapid (>460-fold with our system)<br />

with similar or superior accuracy. Performance <strong>of</strong> fungal classifier at different taxonomic<br />

levels, sequence lengths and regions was also evaluated. Classification<br />

was more accurate with 400-bp sequence reads than with 100-bp reads and location<br />

<strong>of</strong> hyper variable regions played an important role in accuracy <strong>of</strong> sequence<br />

classification. The fungal classifier shows to be a highly effective tool to analyze<br />

large NGS datasets from environmental surveys, and to determine target regions<br />

in LSU and ITS sequences that will improve accuracy <strong>of</strong> taxonomic classification<br />

when short sequences are obtained. The LSU training set and tool are publicly<br />

available through the Ribosomal Database Project (http://rdp.cme.msu.edu/classifier/classifier.jsp)<br />

Porter, Teresita M and G Brian Golding. McMaster University, Biology Department,<br />

Life Sciences Building, 1280 Main Street West, Hamilton, ON L8S<br />

4K1 Canada. Assigning ITS and LSU rDNA ‘barcodes’ with confidence<br />

Whereas PCR, cloning and Sanger-based sequencing can generate libraries<br />

comprised <strong>of</strong> thousands <strong>of</strong> sequences up to about 600-1000 bp in length;<br />

current highly parallel sequencers are now generating amplicon libraries comprised<br />

<strong>of</strong> millions <strong>of</strong> relatively short sequences <strong>of</strong> about 200-400 bp in length. We<br />

are interested in how these shorter sequences affect the quality <strong>of</strong> automated taxonomic<br />

assignments. Here we present the results from two simulation studies that<br />

specifically address the accuracy and error rate <strong>of</strong> ITS and LSU rDNA markers<br />

using several tools suitable for automating assignments. We compare the performance<br />

<strong>of</strong> programs using similarity-, phylogeny-, and composition-based<br />

methods. Generally, we find that database completeness is perhaps the biggest<br />

factor determining assignment accuracy, followed by query sequence length, assignment<br />

method, primer choice, and sequence error. Specifically, we find that<br />

MEGAN lowest common ancestor (LCA) parsing produces the lowest error rate<br />

when assigning unknown ITS and LSU rDNA sequences; and, the fungal LSU<br />

naïve Bayesian classifier available through the Ribosomal Database Project website<br />

performs significantly faster than BLAST-based methods.<br />

Powell, Martha J and Peter M Letcher. Department <strong>of</strong> Biological Sciences, The<br />

University <strong>of</strong> Alabama, Tuscaloosa, AL 35487. Using cellular structure and<br />

biochemistry for insights into fungal cell biology and phylogeny<br />

Chytridiomycota (chytrids) are now recognized as highly adaptable fungi,<br />

not limited in geography or habitat by the constraints <strong>of</strong> reproduction with a flagellated,<br />

unwalled-spore (=zoospore). As our examination <strong>of</strong> chytrid zoospore ultrastructure<br />

has expanded, we have discovered tremendous variety in organellar<br />

structures and organization. Molecular-based reconstructions <strong>of</strong> phylogeny have<br />

also revealed great genetic diversity among chytrids. The intersection <strong>of</strong> molecular<br />

phylogenies with analyses <strong>of</strong> organellar function is helping us understand how<br />

zoospore structures have evolved within lineages. Zoospore architectural differences<br />

are good predictors <strong>of</strong> molecular-based phylogenies. Selective pressures associated<br />

with habitat differences may be drivers in the evolution <strong>of</strong> zoospore ultrastructure.<br />

Evolution <strong>of</strong> zoospore architecture in the Chytridiales is discussed as<br />

an example <strong>of</strong> how structures have been modified and lost within evolutionary lineages.<br />

Our challenge now is discovering more about the functional roles <strong>of</strong> architectural<br />

design differences in chytrid zoospores.<br />

Continued on following page


Presley, Gerry N and Jonathan Schilling. Department <strong>of</strong> Bioproducts and<br />

Biosystems Engineering, University <strong>of</strong> Minnesota, St. Paul, MN 55108. Spatial<br />

mapping <strong>of</strong> gene expression in the brown rot fungus Postia placenta during<br />

wood decay<br />

Wood is an abundant, renewable feedstock for the production bi<strong>of</strong>uels<br />

and biomaterials, however it is very recalcitrant and obtaining fermentable sugars<br />

from it and other lignocellulosic material is energetically costly. Brown rot fungi<br />

are able to obtain carbohydrates from wood at low temperatures by using a unique<br />

mechanism that is thought to involve the production <strong>of</strong> hydroxyl radicals that<br />

modify lignin and allow endoglucanases to access to cellulose in wood cell walls<br />

to aid in the release <strong>of</strong> glucose. Analysis <strong>of</strong> spruce wafers degraded by Postia placenta<br />

suggest that the fungus pretreats its substrate ahead <strong>of</strong> the advancing hyphal<br />

front with oxidative, lignin-modifying chemical reactions followed by saccharification<br />

<strong>of</strong> the substrate, indicated by an increase in endoglucanase activity behind<br />

the hyphal front. This project will focus on determining spatial gene expression<br />

patterns along a spatial gradient in wood degraded by Postia placenta. Wood sections<br />

taken along the length <strong>of</strong> degrading spruce wafers will be analyzed using<br />

Fluorescence in situ hybridization to determine where mRNA from genes thought<br />

to be involved in oxidative pretreatment such as quinone reductases and saccharification<br />

related genes such as endoglucanases are localized in the degrading<br />

wafer. Further analysis <strong>of</strong> gene expression will be done using qRT-PCR to determine<br />

the levels <strong>of</strong> gene expression for pretreatment and saccharification related<br />

genes along the length <strong>of</strong> degrading spruce wafers. This work will spatially map<br />

the expression <strong>of</strong> genes involved brown rot pretreatment and saccharification and<br />

help determine which genes are important in mediating those processes.<br />

Pringle, Anne. Organismic and Evolutionary Biology, Harvard University, Cambridge,<br />

MA, 02138. MSA Student Workshop: Using fungi to get a job<br />

Some ads specifically ask for mycologists, more <strong>of</strong>ten, departments are<br />

looking for conceptually oriented biologists. Fungi <strong>of</strong>fer myriad opportunities for<br />

connecting to broader themes in ecology, evolution, and other fields; making<br />

those connections can be critical to a successful job hunt. I’ll talk about my own<br />

experiences on the job market and share the research and teaching statements I<br />

used to find my current position. We’ll then practice job strategies by taking the<br />

time to talk about our research “mission statements” and “cocktail-party tidbits”.<br />

I’ll explain why these are useful and we’ll finish by sharing our newly crafted mission<br />

statements and tidbits with each other.<br />

Pringle, Anne 1 , Shugeng Cao 2 , and Jon Clardy 2 . 1 Organismic and Evolutionary<br />

Biology, Harvard University, Cambridge, MA, 02138, 2 Harvard Medical School,<br />

Harvard University, Boston, MA, 02115. Chemical diversity across individual<br />

lichens<br />

Fungi <strong>of</strong>ten grow as modular networks, and fungi in the genus Xanthoparmelia<br />

form foliose lichens. Foliose lichens grow radially from a center.<br />

Edges are younger than the middle <strong>of</strong> a thallus, and to test whether modules <strong>of</strong><br />

different ages are chemically equivalent, we took samples from the edges and centers<br />

<strong>of</strong> transects laid across lichens <strong>of</strong> different sizes and used HPLC to characterize<br />

chemical diversity. The modules <strong>of</strong> a single thallus are not equivalent. The<br />

same suite <strong>of</strong> compounds is found across an individual, but dramatic differences<br />

in amounts <strong>of</strong> the different compounds mean the diversity <strong>of</strong> chemicals is greatest<br />

at lichen edges, regardless <strong>of</strong> the size <strong>of</strong> the thallus. The centers <strong>of</strong> very large,<br />

old lichens are chemically depauperate, as compared to edges, but even in small<br />

individuals centers possess fewer amounts <strong>of</strong> compounds as compared to edges,<br />

suggesting the pattern is not driven by age, and is caused instead by the need for<br />

edges to maintain abundant chemical arsenals to colonize new habitats. These<br />

data provide a context for our previous work on bacterial diversity across the thalli<br />

<strong>of</strong> lichens, and are also contextualized by ongoing research on demographies<br />

and life histories <strong>of</strong> filamentous fungi.<br />

Quintanilla, Laura A and Brian D Shaw. Department <strong>of</strong> Plant Pathology and<br />

Microbiology, Program for the Biology <strong>of</strong> Filamentous Fungi, 2132 TAMU,<br />

Texas A&M University, College Station, TX 77843, USA. Actin dynamics in<br />

Aspergillus nidulans<br />

Actin is a major cytoskeletal protein required for the polarized growth <strong>of</strong><br />

filamentous fungi. To study actin dynamics in living cells, Aspergillus nidulans<br />

was transformed with the Lifeact reporter construct that has been used to document<br />

actin patch, cable, and ring dynamics in other systems. Lifeact is a 17 amino<br />

acid peptide derived from the Saccharomyces cerevisiae actin binding protein<br />

Abp140p. Localization patterns for strains expressing Lifeact from three different<br />

promoters displayed similar actin localization patterns and all strains that were<br />

tested grew and developed as wild-type. Lifeact labeled actin localized to a subapical<br />

collar <strong>of</strong> endocytic actin patches that was located approximately 2 µm from<br />

the apex. Cortical actin patches were also present along the periphery <strong>of</strong> the hyphae.<br />

Actin patches displayed anterograde and retrograde movement throughout<br />

the hypha. The presence <strong>of</strong> a dense, complex, and highly dynamic network <strong>of</strong><br />

actin cables (SAW - Sub-apical Actin Web) located at varying distances from the<br />

apex was also observed. This network <strong>of</strong> actin cables was hypothesized to be associated<br />

with branch site or septation site selection. An alternative hypothesis was<br />

that the SAW acted as a diffusion barrier. Actin cables also spanned the length <strong>of</strong><br />

growing hyphae and were associated with the Spitzenkörper. When a hyphal<br />

branch was newly formed, an actin cable network formed inside the parent hypha<br />

and then entered the newly formed branch. A complex <strong>of</strong> actin cables predicted<br />

septation sites and then condensed to form a double acto-myosin ring, which<br />

eventually depolymerized as septal wall material was deposited. Further experiments<br />

designed to test our hypotheses will be discussed.<br />

Ratekin, Angela K 1 , Bernadette C Taylor 1 , and Thomas J Volk 2 . 1 Department<br />

<strong>of</strong> Microbiology, University <strong>of</strong> Wisconsin La Crosse, La Crosse, WI 54601, 2 Department<br />

<strong>of</strong> Biology, University <strong>of</strong> Wisconsin La Crosse, La Crosse, WI 54601.<br />

How bad is BAD1? Looking at the relationship between Blastomyces dermatitidis<br />

Virulence Factor BAD1 and Human Neutrophils, the “Infantry” <strong>of</strong><br />

our Initial Immune Response<br />

Blastomycosis, a fungal infection, presents with a variety <strong>of</strong> disease states.<br />

The causative agent, Blastomyces dermatitidis, a thermally dimorphic fungal<br />

pathogen, is capable <strong>of</strong> infecting immunocompetent hosts. The mechanisms <strong>of</strong> invasion<br />

and the evasion <strong>of</strong> the immune response mounted by the human host are not<br />

clearly understood. The broad based budding yeast form is large, making it difficult<br />

for the phagocytic cells <strong>of</strong> the immune system to engulf. Blastomyces ADhesin<br />

Factor (BAD1) is a virulence factor associated with the yeast form <strong>of</strong> B. dermatitidis.<br />

Neutrophils are the predominant immune cell involved in phagocytosis and<br />

the predominant cell populating the initial defense <strong>of</strong> the immune system. BAD1<br />

attaches to these phagocytic cells resulting in immunomodulation, a change in<br />

chemical messaging. This change may result in an increased activation <strong>of</strong> neutrophils,<br />

resulting in an increase in Neutrophil Extracellur Trap (NET) production.<br />

Since the yeast cells are too large to be readily engulfed by the neutrophils, a net<br />

with a DNA backbone embedded with antimicrobial peptides and enzymes may<br />

aid in the containment and elimination <strong>of</strong> pathogens. However, with some strains<br />

the activation <strong>of</strong> neutrophils may become overzealous resulting in excess lung tissue<br />

damage, presenting as Acute Respiratory Distress Syndrome (ARDS). Looking<br />

at isolates with known clinical presentation, BAD1 expression is measured by<br />

flow cytometry. The production <strong>of</strong> NETs is observed by fluorescence microscopy<br />

and quantified by measuring elastase, a component <strong>of</strong> NETs. A percent killed assay<br />

is being used to determine the survival <strong>of</strong> B. dermatitidis post NET production. Our<br />

hypothesis is that the level <strong>of</strong> BAD1 expression correlates with neutrophil activation<br />

and NET production, which in turn plays a part in determining the type <strong>of</strong> clinical<br />

manifestation. A high level <strong>of</strong> BAD1 expression results in tissue damage,<br />

ARDS, whereas a lower level results in dissemination <strong>of</strong> the organism.<br />

Reynolds, Hannah T 1 , M Kevin Keel 2 , and Hazel A Barton 1 . 1 Department <strong>of</strong><br />

Biology, University <strong>of</strong> Akron, Akron, OH 44325, 2 Department <strong>of</strong> Population<br />

Health, University <strong>of</strong> Georgia, Athens, GA 30602. Assessment <strong>of</strong> the saprotrophic<br />

ability <strong>of</strong> the White Nose Syndrome agent, Geomyces destructans<br />

White Nose Syndrome, an emerging disease <strong>of</strong> hibernating bats, has<br />

killed over 5.5 million bats in northeastern North <strong>America</strong>. Its rapid spread across<br />

the country, beginning in New York in 2006 and currently reaching as far south<br />

as Alabama, has led to numerous cave closures across the United States. The fungus<br />

Geomyces destructans was isolated and described from infected bats and was<br />

recently shown to be the causal agent <strong>of</strong> the disease. Geomyces, the anamorph <strong>of</strong><br />

Pseudogymnoascus, consists <strong>of</strong> 11 species known predominately from cold soils.<br />

Understanding the natural history <strong>of</strong> Geomyces destructans, particularly its ecology<br />

in caves, is crucial for disease management. To investigate the saprotrophic<br />

potential <strong>of</strong> Geomyces destructans, we used a combination <strong>of</strong> enzyme bioassays<br />

and cave community ecology. First, we tested G. destructans and other Geomyces<br />

species for the ability to decompose several common substrates, including chitin,<br />

keratin, and cellulose. We also tested for the production <strong>of</strong> siderophores, ironchelating<br />

molecules that allow survival in low iron conditions and are important<br />

virulence factors for several pathogens. For positive tests, we compared the enzyme<br />

activity <strong>of</strong> G. destructans with that <strong>of</strong> closely related species. Second, we<br />

tested the growth ability <strong>of</strong> G. destructans in sterilized and unsterilized soils and<br />

sediments collected from multiple locations in caves in Cumberland Gap National<br />

Historical Park. For each sample, the microbial communities and soil composition<br />

were analyzed to determine the abiotic and biotic factors that permit the survival<br />

and growth <strong>of</strong> Geomyces destructans in caves.<br />

Reynolds, Nicole K 1 , Emma R Wilson 1 , Prasanna Kandel 1 , Matthew Laramie 1 ,<br />

David Pilliod 2 , and Merlin M White 1 . 1 Boise State University, Department <strong>of</strong> Biological<br />

Sciences, 1910 University Dr. Boise, ID 83725, 2 U.S. Geological Survey,<br />

Snake River Field Station, 970 Lusk St. Boise, ID, 83706. Extending<br />

knowledge <strong>of</strong> Trichomycetes by fixing on preserved immature aquatic insects<br />

in Idaho<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 39


Trichomycetes (gut fungi) are obligate symbionts <strong>of</strong> various arthropods<br />

and have been found in marine, freshwater and terrestrial habitats on every continent<br />

except Antarctica. Minimally, gut fungi associate commensally with their immature<br />

aquatic hosts (including black flies, mayflies, stoneflies, isopods, and others)<br />

attaching to the chitinous lining <strong>of</strong> the mid- or hindgut, although relationships<br />

may shift depending on the situation. Both the geographic distribution and the biodiversity<br />

<strong>of</strong> gut fungi are vastly underestimated. Idaho is no exception, as it presents<br />

many opportunities for discovery in unique habitats, including the sagebrush<br />

steppe. Fourteen locations in remote forest streams <strong>of</strong> Idaho were surveyed for<br />

macroinvertebrates, which were immediately preserved in 95% ethanol upon collection.<br />

Inspection <strong>of</strong> fixed specimens revealed gut fungi were occasionally protruding<br />

beyond the anus <strong>of</strong> the host. Specimens were rehydrated, dissected, slide<br />

mounted and those vouchers <strong>of</strong> gut fungi were used for identification. Hosts included<br />

black flies, mayflies and a few stoneflies. Gut fungi are more typically dissected<br />

from living hosts. These specimens, fixed in the field, presented both challenges<br />

and rewards. Whereas some fungi naturally extend beyond the anus at<br />

maturity, such extension in black fly larvae is not well known. Interestingly, a new<br />

species <strong>of</strong> Genistellospora, which has unusually large tricho- and zygospores, and<br />

is being described from a separate, less remote sampling location, was also observed<br />

in many <strong>of</strong> the black fly samples. We suggest that examination <strong>of</strong> fixed<br />

specimens may present similar opportunities for discovery. Additionally, with<br />

these data collected from fixed specimens, it is clear that these and similar other<br />

habitats in Idaho provide a healthy repository and diversity <strong>of</strong> trichomycetes.<br />

Richards, Thomas A. Dept. <strong>of</strong> Zoology. DC1 406C, The Natural History Museum,<br />

Cromwell Road, London, SW7 5BD, United Kingdom. Horizontal gene<br />

transfer and public goods games in fungi and fungi-like organisms<br />

Horizontal gene transfer (HGT) is the transmission <strong>of</strong> genetic material between<br />

organisms, specifically across species boundaries. A growing body <strong>of</strong> data<br />

suggests that fungi and fungal-like protists have gained genes by HGT. This is an<br />

exciting result because fungi at first glance represent the most recalcitrant <strong>of</strong> all<br />

organisms to gene transfer, possessing robust cell walls and having lost<br />

phagotrophic capacities because they feed exclusively by osmotrophy. Using phylogenetic<br />

methods we investigate the role HGT has played in the evolution <strong>of</strong><br />

fungi and fungal-like protists, including HGT between plants and fungi. This<br />

work demonstrates that HGT has actually played a role in shaping osmotrophic<br />

phenotypes and furthermore has been important for the evolution <strong>of</strong> plant parasitic<br />

mechanisms in the oomycetes. I then use this data to argue that HGT has also<br />

shaped public goods games, an important factor in the evolution <strong>of</strong> osmotrophic<br />

eukaryotes. Together this data suggests HGT in osmotrophic eukaryotes, although<br />

relatively rare compared to prokaryotes, seems to be an important factor for shaping<br />

the evolution <strong>of</strong> biological functions in these groups.<br />

Riley, Rohan 1 , Alexander Idnurm 2 , Philippe Charron 1 , Yolande Dalpé 3 , and<br />

Nicolas Corradi 1 . 1 Canadian Institute for Advanced Research, Department <strong>of</strong> Biology,<br />

2<br />

University <strong>of</strong> Ottawa, 30 Marie Curie Priv. Ottawa ON Canada K1N 6N5,<br />

School <strong>of</strong> Biological Sciences 5007 Rockhill Road University <strong>of</strong> Missouri-<br />

Kansas City Kansas City, MO 64110, USA, 3 Eastern Cereal and Oilseed Research<br />

Center, 960 Carling Ave. Ottawa, Ontario K1A 0C6. Searching for clues<br />

<strong>of</strong> sexual reporoduction in the genomes <strong>of</strong> arbuscular mycorrhizal fungi<br />

Arbuscular Mycorrhizal Fungi (AMF) represent an ecologically important<br />

and evolutionary intriguing group <strong>of</strong> land plant symbionts that have been long<br />

regarded as an ancient asexual lineage. However, the recent acquisition <strong>of</strong> large<br />

sequence datasets has revealed the presence <strong>of</strong> an alternative scenario for their<br />

supposed long evolutionary clonal history; including the presence <strong>of</strong> cryptic sexuality.<br />

Here we report the identification <strong>of</strong> many AMF genes that are commonly<br />

linked with the presence <strong>of</strong> sex in many fungi across the genome <strong>of</strong> different<br />

strains <strong>of</strong> Glomus. These include homologues <strong>of</strong> MATA_HMG proteins, which to<br />

our surprise were found to be extremely diverse in both number and sequence<br />

variant. This elevated diversity is unmatched by any other known fungal species,<br />

and preliminary evidence suggests that some <strong>of</strong> these loci might be involved in<br />

partner recognition in these ecologically important organisms.<br />

Rivas Plata, Eimy, François Lutzoni, Orvo Vitikainen, Trevor Goward, Emmanuel<br />

Sérusiaux, Nicolas Magain, and Jolanta Miadlikowska. Department <strong>of</strong> Biology,<br />

Duke University. 130 Science Drive, Durham, NC 27708-0338. Cophylogenetic<br />

study <strong>of</strong> the lichen-forming fungus Peltigera and its cyanobiont<br />

Nostoc at an intercontinental spatial scale<br />

The Peltigera aphthosa and P. leucophlebia species complexes are classified<br />

within two sections, Peltidea and Chloropeltigera, respectively. These sections<br />

include mostly tripartite lichens associated with a green alga (Coccomyxa) as the<br />

primary photobiont (in the thallus) and a cyanobacterium (Nostoc) as the secondary<br />

photobiont (in the external cephalodia). Only two species, P. malacea and P.<br />

frippii, both from section Peltidea, are bipartite with Nostoc as the only photobiont.<br />

Members <strong>of</strong> both sections are restricted to the coldest regions <strong>of</strong> the northern<br />

hemisphere with a peak <strong>of</strong> species richness in boreal forests. Although P. aph-<br />

40 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

thosa and P. leucophlebia are morphologically similar, to the point <strong>of</strong> being difficult<br />

to distinguish phenotypically, genetically they have diverged extensively<br />

with both clades recovered as paraphyletic but never well supported by bootstrap<br />

analyses. The phenotypical variation in both clades has triggered the description<br />

<strong>of</strong> several species, such as P. britannica, P. chionophila, P. latiloba, and P. nigripunctata.<br />

Using sequence data <strong>of</strong> the internal transcribed spacer (ITS) and two<br />

protein-coding genes (beta-tubulin and RPB1), we inferred phylogenetic relationships<br />

for the fungal partner <strong>of</strong> lichen thalli sampled from North <strong>America</strong>, Europe,<br />

and Asia. Our phylogenies partially support existing species but also suggest a<br />

more complex taxonomic structure that requires a reassessment <strong>of</strong> current species<br />

delimitations in these sections. In parallel, we generated a phylogeny for Nostoc<br />

from the same thalli using the rbcLx locus, and found a pattern <strong>of</strong> associations between<br />

clades <strong>of</strong> lichen mycobionts and Nostoc photobionts. In particular, samples<br />

<strong>of</strong> P. malacea, regardless <strong>of</strong> their geographic origin, appear to be associated with<br />

a unique clade <strong>of</strong> Nostoc, absent in other species, indicating a mycobiont-photobiont<br />

cospeciation event during their coevolution. This rare case <strong>of</strong> cospeciation<br />

could result from a secondary acquisition <strong>of</strong> the bipartite state, following a transition<br />

to a tri-partite state in that section.<br />

Roberson, Robert W. School <strong>of</strong> Life Sciences, Arizona State University, Tempe,<br />

AZ 85287. The structure <strong>of</strong> the hyphal apex: Spitzenkörper or not?<br />

The defining feature <strong>of</strong> filamentous fungi is the hypha. Hyphal growth has<br />

aided the Fungi to successfully utilize a wide range <strong>of</strong> ecological habitats and develop<br />

multiple lifestyles. Cellular and molecular studies <strong>of</strong> hyphal growth have<br />

placed great emphasis on the Spitzenkörper (Spk). The Spk is a dense, roughly<br />

spheroidal cluster <strong>of</strong> vesicles, cytoskeletal components and signaling proteins<br />

found at the tips <strong>of</strong> most growing hyphae. It plays crucial roles in optimizing hyphal<br />

extension rates and in determining patterns <strong>of</strong> growth and morphogenesis by<br />

acting as a dynamic organizing center in the reception <strong>of</strong> secretory vesicles and<br />

orchestrating their delivery to the apical membrane. The Spk appears to have<br />

evolved only in filamentous fungi where it is present in all members <strong>of</strong> the Basidiomycota<br />

and Ascomycota studied thus far. Interestingly, a structural equivalent<br />

<strong>of</strong> the Spk is not common or has not been observed in hyphae <strong>of</strong> the zygomycete<br />

fungi (e.g., Mucoromycotina), or many the other relatively early<br />

diverging lineages. As notable exceptions to this, Spk have been identified in hyphae<br />

<strong>of</strong> Basidiobolus sp. (a zygomycete <strong>of</strong> uncertain phylogeny) and Allomyces<br />

macrogynus (Blastocladiomycota). In fungi that lack a recognizable Spk, vesicle<br />

clusters are <strong>of</strong>ten present in the hyphal apex. Although the arrangement <strong>of</strong> these<br />

vesicles is less organized and complex, there are characteristic patterns that are<br />

recognized. For example, in Rhizopus species, and other members <strong>of</strong> the Mucoromycotina,<br />

a thin crescent-shaped band <strong>of</strong> closely packed vesicles is present<br />

just beneath the apical plasma membrane. Although there have been great advances<br />

in understanding the biology and distribution <strong>of</strong> the Spk, pr<strong>of</strong>ound questions<br />

remain. In this presentation, light and electron microscopy data will used to<br />

review cytoplasmic features <strong>of</strong> the hyphal apex from diverse fungal groups in<br />

hopes to better understand hyphal cell biology and fungal phylogeny.<br />

Robert, Vincent, Joost Stalpers, and Pedro W Crous. CBS-KNAW, Fungal Biodiversity<br />

Center, Uppsalalaan 8, 3534CT Utrecht, The Netherlands. Nomenclatural<br />

databases as working tools for taxonomists: opportunities and challenges<br />

During the 18th IBC held in Melbourne, important decisions were made<br />

to adapt the CODE to accommodate novel developments. One <strong>of</strong> them was to enforce<br />

registration in online repositories. This will guarantee easy access to newly<br />

published taxa and associated data. A number <strong>of</strong> issues remain to be addressed: 1.<br />

possibility <strong>of</strong> multiple repositories that could act as <strong>of</strong>ficial registrars has major<br />

implications. Presently Mycobank (IMA) and Index Fungorum (CABI) are the<br />

only repositories, but more might emerge in the future; 2. coexistence <strong>of</strong> several<br />

repositories necessitates heavy synchronization to reach the desired aims. However,<br />

no collaboration rules are imposed; 3. decisions <strong>of</strong> the NCF are needed with<br />

regards to the data that have to be deposited (minimum and extended datasets); 4.<br />

future scientific developments will require regular adaptations <strong>of</strong> the system (e.g.<br />

environmental sampling, genomic data), with cost implications for repositories; 5.<br />

access to type information <strong>of</strong> taxa should be available and unequivocal to allow<br />

proper taxonomic revisions, especially in the light <strong>of</strong> one fungus one name; 6. registration<br />

<strong>of</strong> species data should be regulated, especially concerning types (epitype,<br />

neotype, lectotype) and barcodes; 7. updates from and relations with journals will<br />

have to be clearly established in order to prevent differences between deposits and<br />

publications; 8. taxonomic data are not always up to date and finding experienced<br />

curators willing to spend time on maintaining taxonomic databases is a challenge<br />

that should not be underestimated; 9. high availability <strong>of</strong> the system has to be ensured<br />

and no serious down-time is acceptable; 10. continuous and pr<strong>of</strong>essional<br />

services will have cost implications with regard to both hard- and s<strong>of</strong>tware. Such<br />

questions will have to be addressed and solved by the mycological community before<br />

January 1, 2013.<br />

Continued on following page


Robinson, Sarah L and Daniel G Panaccione. West Virginia University, Division<br />

<strong>of</strong> Plant and Soil Sciences, Morgantown, WV 26505. Chemotypic and<br />

genotypic diversity in the ergot alkaloid pathway <strong>of</strong> Aspergillus fumigatus<br />

Aspergillus fumigatus is an opportunistic human pathogen that synthesizes<br />

a group <strong>of</strong> mycotoxins via a branch <strong>of</strong> the ergot alkaloid pathway. This fungus<br />

is globally distributed, and genetic data indicate that isolates recombine freely<br />

over that range; however, previous work on ergot alkaloids has focused on a limited<br />

number <strong>of</strong> isolates. We hypothesized that A. fumigatus harbors variation in<br />

the chemotype <strong>of</strong> ergot alkaloids and genotype <strong>of</strong> the ergot alkaloid gene cluster.<br />

Analysis <strong>of</strong> thirteen isolates by high performance liquid chromatography revealed<br />

four distinct ergot alkaloid pr<strong>of</strong>iles or chemotypes. Five isolates completed the A.<br />

fumigatus branch <strong>of</strong> the ergot alkaloid pathway to fumigaclavine C. Six independent<br />

isolates accumulated fumigaclavine A, the pathway intermediate immediately<br />

prior to fumigaclavine C. One isolate accumulated only the early pathway<br />

intermediates chanoclavine-I and chanoclavine-I aldehyde, and one isolate lacked<br />

ergot alkaloids altogether. A genetic basis for each <strong>of</strong> the observed chemotypes<br />

was obtained either by PCR analysis <strong>of</strong> the ergot alkaloid gene cluster or through<br />

sequencing <strong>of</strong> easL, the gene encoding the prenyl transferase that reverse prenylates<br />

fumigaclavine A to fumigaclavine C. Isolates also exhibited differences in<br />

pigmentation and sporulation. The ergot alkaloid chemotypes were widely distributed<br />

geographically and among substrate <strong>of</strong> origin.<br />

Rodriguez Estrada, Alma E 1 , Michael L Draney 1 , and Vicki Medland 2 . 1 University<br />

<strong>of</strong> Wisconsin Green-Bay. Department <strong>of</strong> Natural and Applied Sciences<br />

ES-317. 2420 Nicolet Dr. Green Bay, WI 54311, 2 University <strong>of</strong> Wisconsin<br />

Green-Bay. C<strong>of</strong>rin Center for Biodiversity. Mary Ann C<strong>of</strong>rin Hall 218E. 2420<br />

Nicolet Dr. Green Bay, WI 54311. <strong>Mycological</strong> Research in Panama: A highimpact<br />

field experience for students with variable levels <strong>of</strong> training<br />

Research experiences for undergraduate students have been recently recognized<br />

as a high-impact educational practice to accomplish excellence in any academic<br />

field. The University <strong>of</strong> Wisconsin-Green Bay encourages research experiences<br />

and international courses where undergraduate students have the chance<br />

to fully participate in formal scientific work while experiencing life in a different<br />

culture. The upper level course “Research Experience in Panama” is <strong>of</strong>fered every<br />

year in January. During this course, faculty members and students (biology and<br />

environmental science majors) travel to Panama to perform research for two<br />

weeks. For four years, research themes related to mammals, birds, fish, spiders,<br />

and coral reef and seagrass benthic invertebrates. In 2012, mycological research<br />

was implemented for the first time. This work utilizes a protocol that allows students<br />

with different levels <strong>of</strong> mycological training (undergraduate and masterlevel<br />

students, inexperienced and students who took an upper level mycology<br />

course) to help with ecological macromyetes data collection. This protocol also<br />

allows to rapidly assess the diversity <strong>of</strong> ground-accessible macromycetes using<br />

standardized, repeatable and randomized samples. Data were collected to study<br />

the ecology <strong>of</strong> ectomycorrhizal and saprotrophic macr<strong>of</strong>ungi at a variety <strong>of</strong> sites<br />

near Bocas del Toro and Gamboa field stations (Smithsonian Tropical Research<br />

Institute). Specifically, this research seeks to study the effects <strong>of</strong> altitude and vegetation<br />

on macromycetes richness and abundance across a range <strong>of</strong> Panamanian<br />

rainforest (lowland, mid and high elevation). Parallel transects <strong>of</strong> 250 m 2 (50 x 5<br />

m) were delimited within 0.25 hectare sites where macromycetes were collected,<br />

described and counted. The sites sampled in 2012 will be sampled in subsequent<br />

years, and different students will participate. Ecological research on<br />

macromycetes is a high-impact practice that strengths students’ understanding <strong>of</strong><br />

science and elevates students’ awareness <strong>of</strong> the importance <strong>of</strong> fungi in the ecosystems.<br />

Rosales, Antonio 1 , Terri Tobias 1 , Katherine Suding 2 , Robert Sinsabaugh 3 , and<br />

Andrea Porras-Alfaro 1 . 1 Department <strong>of</strong> Biological Sciences, Western Illinois<br />

University, 2 Department <strong>of</strong> Environmental Sciences, University <strong>of</strong> California,<br />

Berkeley, 3 Department <strong>of</strong> Biology, University <strong>of</strong> New Mexico. Vertically transmitted<br />

endophytes from plants in the alpine tundra<br />

Alpine tundra plants are able to survive harsh conditions with the help <strong>of</strong><br />

endophytes and mycorrhizal fungi. Endophytes are fungi or bacteria that live<br />

within plants without causing apparent damage. These symbiotic relationships between<br />

plants and endophytic fungi are known to have a direct impact on plant<br />

community structure, fitness, and diversity. Endophytes can colonize plants by either<br />

horizontal or vertical transmission and little is known about their abundance,<br />

taxonomy and function. The objective <strong>of</strong> this study was to isolate and quantify<br />

vertically transmitted endophytes from plant seeds that were collected at the<br />

Niwot LTER Site in Colorado. Six plant species (Geum rossii, Erigeron simplex,<br />

Artemisia scopulorum, Deschampsia cespitosa, Bistorta bistortoides, and Trisetum<br />

spicatum) were studied. Thirty seeds <strong>of</strong> each plant were surface sterilized and<br />

plated on malt extract agar with antibiotics. Fungi were sequenced using the ITS<br />

rDNA region. Thirty-three endophytes were isolated from the seeds. Ten different<br />

morphotypes were present. E. simplex and D. cespitosa show the highest ger-<br />

mination rates. Germination rates varied between 67% to 0% and colonization<br />

rates varied between 6% to 33%. B. bistortoides seeds did not germinate and<br />

showed the highest fungal colonization rates. Preliminary identification <strong>of</strong> some<br />

isolates using ITS rDNA shows that G. rossii and D. cespitosa seeds are colonized<br />

by a fungus closely related to Cladosporium uredinicola. Additional tests need to<br />

be conducted to determine the potential role that these seed endophytes may have<br />

in plant germination, seed establishment and the structure <strong>of</strong> plant communities in<br />

the alpine tundra.<br />

Runa, Farhana 1 , Ignazio Carbone 1 , Deepak Bhatnager 2 , and Gary A Payne 1 1<br />

.<br />

2<br />

Department <strong>of</strong> Plant pathology, North Carolina State University, Raleigh, NC,<br />

Southern Regional Research Center, USDA , New Orleans, LA. Nuclear condition<br />

<strong>of</strong> Aspergillus flavus by heterokaryosis and ploidy analysis<br />

Conidia <strong>of</strong> Aspergillus flavus are multinucleate but it is unknown whether<br />

they are predominantly homo- or heterokaryotic. A heterokaryon could serve as a<br />

source <strong>of</strong> genetic diversity within strains for pathogenicity and aflatoxin production<br />

if homokaryotic as well as heterokaryotic nuclei are packaged selectively into<br />

conidia during conidiation. To study nuclear condition in A. flavus we transformed<br />

strain AFC (arg7, pyrG) with either plasmid HH2B-ECFP to create AFC-<br />

ECFP (pyr), a strain expressing nuclei with cyan fluoresce, or with plasmid<br />

HH2A-EYFP to create AFC-EYFP (arg), a strain expressing yellow fluorescing<br />

nuclei. Protoplasts from the two strains were subjected to polyethylene glycol mediated<br />

cell fusion. Putative fusants were selected for their ability to grow on minimal<br />

medium, which should not allow the growth <strong>of</strong> neither nutritional mutant<br />

AFC-ECFP (pyr) nor AFC-EYFP (arg). The putative fusants produced mycelia<br />

with some nuclei expressing ECFP and other nuclei expressing EYFP. We also<br />

observed that conidia from these fusants have three types <strong>of</strong> nuclei: only EYFP<br />

expressing, only ECFP expressing, or both EYFP and ECFP expressing. Ploidy<br />

analysis by flow cytometry indicated that the putative 11-2 fusant is heterokaryotic.<br />

Moreover, flow cytometry and merged fluorescence microscopy image data<br />

showed that the majority <strong>of</strong> conidia within 11-2 are homokaryotic, expressing either<br />

EYFP or ECFP in subsequent generations; however, a very small percentage<br />

<strong>of</strong> conidia maintain nuclei expressing both. Conidia having nuclei with merged<br />

fluorescence have been separated and sorted by FACS. We will determine<br />

whether the merged fluorescence microscopy images are indicative <strong>of</strong> nuclei expressing<br />

EYFP+ECFP, maintained predominantly as heterokaryons or as<br />

diploids. Confocal microscopy will be applied to track the migration <strong>of</strong><br />

EYFP+ECFP nuclei during germination and conidiation. This work will allow us<br />

to understand the mechanism whereby heterokaryons or diploids are maintained<br />

in multinucleated conidia <strong>of</strong> A. flavus.<br />

Ryan, Katy L and Daniel G Panaccione. West Virginia University, Division <strong>of</strong><br />

Plant and Soil Sciences, Morgantown, WV 26505. Partial reconstruction <strong>of</strong> the<br />

ergot alkaloid pathway in Aspergillus nidulans<br />

Ergot alkaloids are pharmaceutically and agriculturally important secondary<br />

metabolites produced by several species <strong>of</strong> fungi. A better understanding <strong>of</strong><br />

the biosynthetic pathway <strong>of</strong> ergot alkaloids may allow for modification <strong>of</strong> these<br />

compounds to produce effective pharmaceuticals or control their accumulation in<br />

agriculture. Ergot alkaloid pathways vary among fungi, but the pathway intermediate<br />

chanoclavine appears to be evolutionarily conserved among ergot alkaloid<br />

producers. Chanoclavine synthesis occurs in at least five steps, and knockout and<br />

heterologous expression studies have demonstrated that four genes, dmaW, easF,<br />

easE, and easC, are required for pathway steps prior to chanoclavine. However,<br />

it is currently unknown whether these genes are sufficient to synthesize chanoclavine.<br />

DmaW prenylates tryptophan to produce dimethylallyltryptophan<br />

(DMAT), and easF encodes a methyl transferase that produces the second intermediate,<br />

N-Me-DMAT. The roles <strong>of</strong> EasE and EasC have not been established.<br />

Four genes, dmaW, easF, easE, and easC were amplified from the human<br />

pathogen Aspergillus fumigatus and transformed into Aspergillus nidulans, which<br />

does not contain any <strong>of</strong> the ergot alkaloid synthesis genes and produces no ergot<br />

alkaloids. We determined by HPLC and LC/MS that the A. nidulans strains containing<br />

the four A. fumigatus genes produced chanoclavine and earlier pathway intermediates.<br />

The specific roles <strong>of</strong> EasE and EasC are being investigated through<br />

the generation <strong>of</strong> transformants containing DmaW, EasF, and EasC, or DmaW,<br />

EasF, and EasE, compared to appropriate controls. HPLC and LC/MS data<br />

demonstrated that EasC catalyzes the formation <strong>of</strong> a new intermediate; the role <strong>of</strong><br />

EasE is still being investigated. We conclude that the four genes listed above are<br />

sufficient for the synthesis <strong>of</strong> chanoclavine in A. nidulans and that our approach<br />

<strong>of</strong> expressing ergot pathway genes in A. nidulans will provide a better understanding<br />

<strong>of</strong> the early steps in ergot alkaloid synthesis.<br />

Ryu, Jae San, Kyung Hee Kim, Min Keun Kim, Su Myung Chae, Hee Jeong Je,<br />

and Won Du Song. Eco-friendliness Research Department, Gyeongsangnam-do<br />

Agricultural Research and Extension Services, Jinju 660-360, Republic <strong>of</strong> Korea.<br />

The genetic structure <strong>of</strong> mating type locus B3 in Pleurotus eryngii<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 41


King oyster mushroom Pleurotus eryngii is a heterothallic homobasidiomycete<br />

with a tetrapolar incompatibility system controlled by the two mating<br />

type loci, A and B. RAPD and BSA were performed to screen primers linked to<br />

the specific mating types. Out <strong>of</strong> 100 random primers, OPS-03 showed the specific<br />

band, OPS-032100 co-segregated with AxB3. The mating type B3 locus was<br />

identified based on a SCAR marker derived from OPS-032100 from a genomic<br />

library <strong>of</strong> the dikaryotic strain KNR2312. Sequence analysis suggested the genetic<br />

basis for B3 locus in P. eryngii was similar to those in model mushrooms. Interestingly,<br />

4 putative pheromone receptor genes, instead <strong>of</strong> 3 that are typical, and<br />

6 putative pheromone genes were predicted. Every pheromone receptor was<br />

packed with two pheromone precursors in three receptors while the extra one receptor;<br />

pepr3 does not have any cognate pheromone beside its sequences. B3<br />

locus specific RNAs were isolated and reverse transcribed further into cDNA to<br />

characterize their sequence and role as signal factors. qRT-PCR showed transcription<br />

levels <strong>of</strong> the pheromone and the pheromone receptor in monokaryons<br />

were much higher than in a dikaryon. Such a study could be useful for developing<br />

self-fertilizing monokaryotic strains using RNAi (knock-down) and for deciphering<br />

comparative gene expression.<br />

Ryu, Jae San 1 , Song Hee Lee 2 , Kyung Hee Kim 1 , Hee Jeong Je 1 , Min Keun<br />

Kim 1 , Hyun sook Lee 2 , and Su Myung Chae 1 . 1 Eco-friendliness Research Department,<br />

Gyeongsangnam-do Agricultural Research and Extension Services,<br />

Jinju 660-360, Republic <strong>of</strong> Korea, 2 Departments <strong>of</strong> Microbiology, Gyeongsang<br />

National University, Jinju, 660-900, Republic <strong>of</strong> Korea. Complete genome sequence<br />

<strong>of</strong> Pleurotus eryngii using the next generation sequencing<br />

Pleurotus eryngii (King oyster mushroom) is an edible mushroom with potential<br />

health benefits that thrives on a parasitic host in the steppe area. P. eryngii<br />

is a basidiomycete with a tetrapolar incompatibility system comprising two mating<br />

type loci, A and B. Two monokaryons P5 and P6 from P. eryngii KNR 2312 strain<br />

by a de-dikaryotization were analyzed by CHEF, revealing 10 chromosomal bands<br />

with a size range from 3.0 to 7.1 Mbp and an estimated total genome size <strong>of</strong> 43.8<br />

Mbp. GS-FLX and HiSeq TM Sequencing analysis <strong>of</strong> P5 as a Next Generation Sequencing<br />

(NGS) strategy was performed to clarify its genetic composition. Shotgun<br />

and mate paired libraries were constructed for NGS. The read count was<br />

650,812 and 433,267,606 total bases for the shotgun library. Average length <strong>of</strong><br />

read was over 665 bases. For the mate paired library, read count was 122,368, total<br />

bases was 47,889,870, and average <strong>of</strong> length was 391.36. Scaffolding was performed<br />

to assemble contigs and mate-paired data. The number <strong>of</strong> scaffolds and<br />

bases were 1,506 and 43,991,976, respectively. The scaffold accounted for 100%<br />

<strong>of</strong> the genome size estimated by karyotyping. To estimate function, sequences<br />

were compared with those in the NCBI blastx database. Over 1,000 genes were<br />

aligned with the sequences within the contigs. Functional genomics will be required<br />

to elucidate unknown genes. Results <strong>of</strong> this genome project will soon allow<br />

sequence-based breeding like MAS (marker assisted selection).<br />

Sadowsky, Jesse J, Linda TA van Diepen, and Serita D Frey. University <strong>of</strong> New<br />

Hampshire, 114 James Hall, Durham, NH 03824. Contributions <strong>of</strong> ectomycorrhizal<br />

fungi to organic matter formation and degradation in response to<br />

chronic nitrogen deposition<br />

Nitrogen deposition promotes carbon storage in temperate forests by altering<br />

organic matter decomposition processes and decomposer communities. Ectomycorrhizal<br />

(ECM) fungi participate in soil organic matter dynamics by producing<br />

abundant mycelium and actively degrading labile and recalcitrant organic<br />

matter, but have not been considered explicitly for their role in organic matter contribution<br />

and decomposition in simulated long-term nitrogen enrichment studies.<br />

Plots within a temperate mixed-hardwood stand at the Harvard Forest Long-Term<br />

Ecological Research site in Petersham, MA, USA have been amended with 50 or<br />

150 kg N ha -1 yr -1 since 1988 to simulate projected future N deposition levels or<br />

N saturation, respectively. We determined how chronic N deposition affects annual<br />

organic matter input by ECM fungi by collecting hyphae in root-excluded<br />

sand bags. We assessed potential organic matter decomposition by ECM fungal<br />

communities by measuring activity <strong>of</strong> five C, N, or P hydrolases and two oxidases<br />

on the mantle <strong>of</strong> ECM root tips and <strong>of</strong> foraging extraradical mycelium collected<br />

in sand bags. Fungal symbionts were identified by DNA sequencing. ECM hyphal<br />

production was reduced at the highest nitrogen deposition rate. Nitrogen<br />

deposition increased cellulolytic and chitinolytic activity <strong>of</strong> ECM root tips in the<br />

organic soil horizon and decreased aminopeptidase activity in the mineral soil<br />

horizon. This generally agrees with earlier observations in bulk soil at this and<br />

other sites receiving multiple years <strong>of</strong> nitrogen enrichment. In root-excluded sand,<br />

chitinase activity increased by three-fold at both rates <strong>of</strong> added nitrogen, while<br />

laccase activity declined by one-third and three-fourths at low and high nitrogen<br />

rates. ECM fungal species richness was least at the highest nitrogen deposition<br />

rate. Further efforts towards determining the fate <strong>of</strong> organic matter pools contributed<br />

and decomposed by ECM fungi communities will help to clarify their<br />

role in soil organic matter dynamics under elevated nitrogen deposition and other<br />

anthropogenic disturbances.<br />

42 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Salamone, Amy and Allison Walker. Department <strong>of</strong> Coastal Sciences, University<br />

<strong>of</strong> Southern Mississippi, 703 East Beach Dr., Ocean Springs, MS 39564. On<br />

the rocks: the first look at fungal biodiversity in marine bi<strong>of</strong>ilms <strong>of</strong> submerged<br />

artificial reefs<br />

This study was the first characterization <strong>of</strong> natural mixed-species fungal<br />

communities in bi<strong>of</strong>ilm on artificial reefs in the marine environment. Previous<br />

fungal bi<strong>of</strong>ilm studies have failed to observe the naturally occurring biodiversity<br />

in coastal areas, which are the most utilized zones <strong>of</strong> the marine environment.<br />

Fungi have evolved to form bi<strong>of</strong>ilms for protection, to communicate via quorum<br />

sensing, to gain access to nutrients, and to exchange genetic information. These<br />

communities fuel many primary consumers that support artificial reef habitats.<br />

Fungi, along with bacteria, archaea, protists, algae, and diatoms, quickly form a<br />

complex bi<strong>of</strong>ilm on immersed surfaces in seawater. This settlement and succession<br />

calls for further examination, given the major trophic implications therein, as<br />

properties <strong>of</strong> the bi<strong>of</strong>ilm can influence future establishment <strong>of</strong> other marine organisms.<br />

For this study, an optimized extraction technique for fungal DNA from<br />

bi<strong>of</strong>ilm was developed. Fungal bi<strong>of</strong>ilm communities were characterized by culturing,<br />

morphological identification, ITS T-RFLP analysis, and ITS gene sequencing.<br />

The role <strong>of</strong> seasonality and other abiotic factors in structuring these<br />

communities was also examined. Fungal biodiversity in these bi<strong>of</strong>ilms is high,<br />

spanning over 25 genera, including species <strong>of</strong> Ascomycetes, Basidiomycetes, Zygomycetes,<br />

and potential novel yeast species. This assessment <strong>of</strong> fungal community<br />

biodiversity in marine bi<strong>of</strong>ilm yields insight into artificial reef fungal succession<br />

and functionality in the North-Central Gulf <strong>of</strong> Mexico.<br />

Salgado-Salazar, Catalina 1 , Amy Y Rossman 2 , and Priscila Chaverri 1 . 1 2102<br />

Plant Sciences Bld, University <strong>of</strong> Maryland, College Park, MD 20742, 2 USDA-<br />

ARS, Systematic Mycology and Microbiology Laboratory, Beltsville, Maryland<br />

20705. Molecular systematics <strong>of</strong> saprophytic and plant pathogenic isolates <strong>of</strong><br />

the cosmopolitan fungus Thelonectria discophora (Nectriaceae, Hypocreales,<br />

Ascomycota)<br />

Thelonectria discophora, previously placed in Neonectria, (anamorph<br />

“Cylindrocarpon” ianthothele var. majus) is a cosmopolitan species found on a<br />

wide range <strong>of</strong> living or recently dead woody plants, including gymnosperms and<br />

dicotyledonous plants and less frequently on palm trunks and herbaceous materials.<br />

A variety <strong>of</strong> this species, “Neonectria” discophora var. rubi, has been associated<br />

with a distinctive basal canker <strong>of</strong> cultivated Rubus sp. Molecular techniques<br />

have been useful to define species limits and have revealed that many widespread<br />

and cosmopolitan organisms can be formed by multiple genetically distinct cryptic<br />

species or sub-species that may be geographically or ecologically correlated.<br />

In order to test if T. discophora is a group <strong>of</strong> cryptic species with geographic or<br />

ecological restrictions, multi-gene phylogenetic analyses were conducted using<br />

Maximum Likelihood and Bayesian Inference approaches on a worldwide collection<br />

<strong>of</strong> isolates. Fifteen cryptic groups could be distinguished within T. discophora<br />

distributed in three major clades. These clades were supported by high<br />

bootstrap values and posterior probabilities. Some cryptic groups were composed<br />

by isolates coming from close geographic locations; however geographic clustering<br />

is not the rule as isolates from distant regions also grouped together. Interestingly,<br />

saprophytic isolates on plants, soil and pathogens were found to form separate<br />

groups seemingly isolated by ecology. Genetic divergence and fixation<br />

index values indicate that the cryptic groups are highly divergent, not forming a<br />

single panmictic population.<br />

Sanchez-Garcia, Marisol and P Brandon Matheny. Department <strong>of</strong> Ecology and<br />

Evolutionary Biology, University <strong>of</strong> Tennessee, Knoxville, TN, 37996-1610,<br />

USA. A phylogenetic evaluation <strong>of</strong> the tribe Leucopaxilleae: polyphyly, the<br />

LPD grade, and novel taxa from the Southern Appalachians<br />

The tribe Leucopaxilleae Singer comprises eight genera that have been<br />

placed at various taxonomic ranks. Phylogenetic analysis <strong>of</strong> a four-gene region<br />

supermatrix (nuclear rRNA regions, rpb2) shows the tribe to be highly polyphyletic,<br />

with genera distributed in the Tricholomatoid, Marasmioid, Hygrophoroid<br />

and Pluteoid clades. Leucopaxillus, Porpoloma and Dennisiomyces,<br />

occur in the Tricholomatoid clade. Dennisiomyces, primarily a neotropical genus,<br />

has not been previously considered part <strong>of</strong> the Leucopaxilleae. We present preliminary<br />

data that suggest Dennisiomyces is closely related to Leucopaxillus and<br />

Porpoloma forming the ‘LPD’ grade from which the ectomycorrhizal (ECM)<br />

genus Tricholoma appears to be derived. These results also suggest that Leucopaxillus<br />

and Porpoloma are polyphyletic. Additionally, three unknown species<br />

that form a clade nested in the ‘LPD’ grade have been collected recently from the<br />

Southern Appalachians <strong>of</strong> the southeast United States. They are distinguished<br />

from other species in the LPD grade by possession <strong>of</strong> distinct cheilocystidia and<br />

pleurocystidia, presence <strong>of</strong> clamp connections and inamyloid, smooth spores. At<br />

present, we are unable to ascribe any known genus to this clade.<br />

Continued on following page


Sandberg, Dustin C and A Elizabeth Arnold. School <strong>of</strong> Plant Sciences, The University<br />

<strong>of</strong> Arizona, Tucson, AZ 85721 USA. Host- and geographic structure <strong>of</strong><br />

endophyte communities in aquatic plants <strong>of</strong> northern Arizona<br />

Many <strong>of</strong> the general themes in endophyte biology have been developed<br />

from studies <strong>of</strong> terrestrial plants. However, aquatic plants represent a suite <strong>of</strong> morphological<br />

and physical traits that are distinct from those living on land: thin cuticles,<br />

frequently open stomata, and specialized roots that are <strong>of</strong>ten adapted for<br />

oxygen uptake. The goal <strong>of</strong> this study was to evaluate the diversity, composition,<br />

host- and tissue affiliations, and geographic structure <strong>of</strong> fungal endophytes associated<br />

with diverse aquatic plants. In Fall 2011, the four most abundant species <strong>of</strong><br />

submergent and emergent macrophytes were collected from three sites in each <strong>of</strong><br />

six reservoirs and lakes in north-central Arizona. Endophytes were isolated in culture<br />

from roots, stems, and photosynthetic tissues <strong>of</strong> each species, and were evaluated<br />

using a culture-free approach by cloning directly from plant genomic DNA<br />

isolated from each tissue type. All isolates and clones were sequenced for a<br />

1000bp fragment comprising the internal transcribed spacers, 5.8S gene, and partial<br />

nuclear ribosomal large subunit. The resulting data were used to evaluate five<br />

predictions: (1) endophytes <strong>of</strong> aquatic plants will be dominated by cosmopolitan<br />

genera <strong>of</strong> fungi that can take advantage <strong>of</strong> the limited structural defenses <strong>of</strong> aquatic<br />

plants to colonize them more readily than terrestrial plants; (2) endophytes <strong>of</strong><br />

aquatic plants will be diverse, with many different lineages capable <strong>of</strong> persisting<br />

in tissues <strong>of</strong> macrophytes; (3) endophyte communities will differ as a function <strong>of</strong><br />

tissue type and depth beneath the water surface; (4) endophyte communities will<br />

be similar among hosts within lakes but will differ between lakes; (5) culture-free<br />

approach will reveal more specialized endophytes, including aquatic specialists,<br />

than culture-based surveys. These analyses provide a first quantitative estimation<br />

<strong>of</strong> endophytic associations in diverse, ecologically important, and economically<br />

relevant aquatic plants, and provide a basis for evaluating general trends based on<br />

plants in terrestrial ecosystems.<br />

Sandona, Katrina 1 , Terri Tobias 1 , Antonio Rosales 1 , Katherine Suding 2 ,<br />

Robert Sinsabaugh 3 , and Andrea Porras-Alfaro 1 . 1 Western Illinois University, 1<br />

University Circle, Macomb IL. 61455, 2 University <strong>of</strong> California Berkeley, 130<br />

Mulford Hall 3144, Berkeley CA. 94720, 3 University <strong>of</strong> New Mexico, 167A<br />

Castetter Hall, Albuquerque NM. 87131. Effect <strong>of</strong> a dominant endophytic fungus,<br />

Phialocephala fortinii, on plant growth<br />

Dark septate endophytes have been detected in over 600 plant species<br />

worldwide. Phialocephala fortinii have been reported multiple times as a common<br />

dark septate endophyte in the alpine tundra, heathlands, and forests. The objective<br />

<strong>of</strong> this study was to determine the effect on seed germination and plant<br />

growth <strong>of</strong> Phialocephala fortinii isolates on different commercial plants. Fungi<br />

were isolated from roots that were collected at the Niwot Long Term Ecological<br />

Research Site in Colorado. Root isolates were identified using ITS rDNA and<br />

Phialocephala fortinii was selected for germination experiments. The fungus<br />

SS37 was plated in five culture jars containing malt extract agar with antibiotics.<br />

Five jars with no fungus were used as controls. After a week six surfaced sterilized<br />

seeds <strong>of</strong> Zea mays (corn) and Glycine max (soybean) where planted in each<br />

jar. Seeds were allowed to grow for one week. Plants were harvested from the<br />

jars; the roots and stems were measured. The fungus Phialocephala significantly<br />

stimulated the growth <strong>of</strong> corn seeds with an average number <strong>of</strong> roots <strong>of</strong> 8.3±0.3<br />

(SE) with respect to 2.4±0.3 SE for the control (P


merization <strong>of</strong> wood with little weight loss, using a reduction-oxidation pathway<br />

that yields highly reactive hydroxyl radicals through a chelator-mediated Fenton<br />

reaction (CMFR). While hydroxyl radicals would readily damage cellulases, the<br />

fungus is believed to conduct both reactions, enzymatic and oxidative, in the same<br />

wood sample. We are examining the spatial and temporal relationship <strong>of</strong> these<br />

two reaction systems, complemented by pro<strong>of</strong>-<strong>of</strong>-concept biomimick experiments.<br />

The results have shown clear reaction partitioning at coarse scale in wood<br />

wafers, which will be presented. The outcomes from such fundamental inquiry in<br />

a novel application context are useful, given the traditional pest-oriented focus on<br />

these fungi and the targeted focus on wood strength loss. Our systems-based research<br />

should help better understand these fungi, not only for potential in application<br />

but as key players in the global carbon cycle.<br />

Seifert, Keith A 1 and Andrew N Miller 2 . 1 Eastern Cereal & Oilseed Research<br />

Centre,<br />

2<br />

Agriculture & Agri-Food Canada, Ottawa, Ontario, Canada K1A 0C6,<br />

University <strong>of</strong> Illinois, Illinois Natural History Survey, Champaign, IL 61820-<br />

6970 USA. Subcommissions, working groups, and protected lists<br />

The International Commission on the Taxonomy <strong>of</strong> Fungi (ICTF) is a 20member<br />

group affiliated with both the International <strong>Mycological</strong> Association and<br />

the International Union <strong>of</strong> Microbiological Societies. Its activities over the past 30<br />

years have been focused on the taxonomy and nomenclature <strong>of</strong> economically important<br />

fungal groups, with subcommissions or affiliated commissions working<br />

on Penicillium and Aspergillus, Trichoderma and Hypocrea, and Fusarium. Informal<br />

working groups on Ceratocystis and Ophiostoma, and on the Mycosphaerella<br />

complex, have been periodically active. In the wake <strong>of</strong> the nomenclature<br />

changes discussed in this symposium, additional groups on Colletotrichum<br />

and the Hypocreales are in the process <strong>of</strong> forming. The activities <strong>of</strong> these groups<br />

will briefly be reviewed in this talk. Although over the long term the ICTF will<br />

continue its focus on economically important fungi, in the short term we plan to<br />

assist with the formation <strong>of</strong> working groups to address the broadest possible taxonomic<br />

scope. This is an exciting time for increased cooperation and collaboration<br />

among fungal taxonomists, as we work towards developing consensus on the<br />

selection <strong>of</strong> genus and species names, and the development <strong>of</strong> lists <strong>of</strong> protected<br />

and/or rejected names.<br />

Shaw, Brian D, Da-Woon Chung, Srijana Upadhyay, Charles Greenwald,<br />

Sheng-Li Deng, Heather H Wilkinson, and Daniel J Ebbole. Program for the Biology<br />

<strong>of</strong> Filamentous Fungi, Department <strong>of</strong> Plant Pathology and Microbiology,<br />

2132 TAMU, Texas A&M University, College Station TX 77843, USA. Regulation<br />

<strong>of</strong> conidiation in Aspergillus nidulans and Neurospora crassa<br />

Did conidiation arise more than once in ascomycete lineages or did extant<br />

condiation strategies diverge from an ancestral form? To begin to address these<br />

questions, transcriptional regulation <strong>of</strong> conidiation from the two best-studied conidiating<br />

ascomycetes, Aspergillus nidulans and Neurospora crassa, was examined.<br />

We hypothesized that a conidiation regulatory pathway was present in the<br />

ancestral species, and became specialized in the extant species by gaining and/or<br />

losing specific regulators. N. crassa orthologs <strong>of</strong> seven regulatory genes in A.<br />

nidulans (fluG, flbC, flbD, abaA, wetA, medA, and stuA) were examined. Expression<br />

<strong>of</strong> the N. crassa orthologs complemented defective conidiation in the A. nidulans<br />

fluG, flbD, wetA, medA, and stuA mutants. We detected four patterns for the<br />

conidiation regulators: (i) Non-homologous genes with analogous roles in conidiation<br />

(brlA and fl), (ii) Orthologs with retained biochemical function that lack an<br />

analogous role in conidiation (fluG, flbD, and wetA), (iii) Orthologs with retained<br />

biochemical function with analogous roles in conidiation (medA and stuA), and<br />

(iv) Orthologs with biochemical function not conserved with analogous roles in<br />

conidiation (abaA). It is reasonable to expect that if the extant conidiation strategies<br />

for these two species evolved from the same ancestral pathway the sets <strong>of</strong><br />

genes co-regulated with known targets <strong>of</strong> conidiation regulators would exhibit<br />

considerable overlap. However, transcriptional pr<strong>of</strong>iling <strong>of</strong> both species across a<br />

time course <strong>of</strong> conidiation using RNASeq revealed no overlap in co-regulaiton <strong>of</strong><br />

these gene sets, thus providing no evidence for shared mechanisms below the regulators.<br />

Taken together, our data for the roles <strong>of</strong> conidiation regulator orthologs<br />

and the behavior <strong>of</strong> known conidiation associated genes that are likely targets <strong>of</strong><br />

regulators across these distantly related species provides little support for a common<br />

pathway.<br />

Shaw, Jeffrey, Daniel Spakowicz, Cambria Alpha, Brian Dunican, Kaury<br />

Kucera, and Scott Strobel. Department <strong>of</strong> Molecular Biophysics and Biochemistry,<br />

Yale University, 260 Whitney Ave., JW Gibbs Room 321, New Haven, CT<br />

06520. Biological and volatile organic compound diversity in fungal endophytes<br />

from Malaysian and Ecuadorian rainforests<br />

Fungal endophytes comprise a ubiquitous yet understudied source <strong>of</strong> biological<br />

diversity. We have targeted this endophytic diversity for study as part <strong>of</strong><br />

an effort to identify new secondary metabolite chemistry in microorganisms. The<br />

concentration <strong>of</strong> endophyte diversity has been shown be greatest in tropical rainforests<br />

so, through collaborations with the governments <strong>of</strong> Ecuador and the<br />

44 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Malaysian state <strong>of</strong> Sarawak, we have undertaken collection expeditions to investigate<br />

the biological and chemical diversity <strong>of</strong> endophytes in these regions. The<br />

Ecuador expedition is done in conjunction with the Yale “Rainforest Expedition<br />

and Laboratory” course which includes 15 undergraduate students. Based on sequencing<br />

<strong>of</strong> the internal transcribed spacer (ITS) region <strong>of</strong> the ribosomal RNA in<br />

our isolates and comparison to GenBank, the endophytes isolated from the rainforests<br />

<strong>of</strong> Ecuador and Malaysia are among the most divergent fungi currently<br />

known. Of the isolates collected, 23% <strong>of</strong> those from Sarawak (36 total) and 18%<br />

<strong>of</strong> those from Ecuador (121 total) have less than 95% identity to any cultured organism.<br />

This metric can be used to roughly estimate the number <strong>of</strong> new species<br />

in our collection, although ITS divergence among species <strong>of</strong> fungi can vary widely.<br />

Of the total sample, 72 isolates (approximately 9%) have less than 85% identity<br />

to the cultured organisms in GenBank. Isolates from these expeditions were<br />

assayed for volatile organic compound production (VOC) by GC/MS. We have<br />

identified several unique metabolites produced by these organisms that have not<br />

been previously observed. Among these are a series <strong>of</strong> molecules, produced by<br />

several different fungi, identified as nonatetraenes. Other hydrocarbons were produced<br />

with a range <strong>of</strong> carbon chain lengths (C8 - C20 ) and varying degrees <strong>of</strong> unsaturation,<br />

branching and cyclization. Several <strong>of</strong> the compounds identified have<br />

interesting bioactivity or properties that may make them attractive bi<strong>of</strong>uels.<br />

Shu, Xiaomei. Department <strong>of</strong> Plant Pathology, Box 7567, 267 Partners III, North<br />

Carolina State University, Raleigh, NC 27695-7567. Pathogenesis and host response<br />

during infection <strong>of</strong> seeds by Aspergillus flavus and Fusarium verticillioides<br />

Both Aspergillus flavus, a saprophyte, and Fusarium verticillioides, an endophyte,<br />

can colonize developing maize seeds and contaminate them with mycotoxins.<br />

Susceptibility is conditioned by plant stress, but little is known about the<br />

infection processes by the two fungi, and whether their different trophic lifestyles<br />

affect tissue colonization and host response. To unravel the biology and pathogenesis<br />

<strong>of</strong> these two fungi, we followed fungal infection and host defense response<br />

<strong>of</strong> developing maize seed for three days after introducing the conidial suspension<br />

through the seed pericarp. Histological examination <strong>of</strong> thin sections <strong>of</strong><br />

maize seeds showed that the fungi were restricted to the site <strong>of</strong> inoculation 48<br />

hours post inoculation. Within the next 24 hours mycelium was detected in the<br />

aleurone, the endosperm and the embryo-endosperm interface. Mycelium <strong>of</strong> A.<br />

flavus was observed on the surface <strong>of</strong> either enlarged or shrunken cells <strong>of</strong> the aleurone,<br />

but never inside intact cells. In contrast, F. verticillioides was sometimes observed<br />

inside intact aleurone cells. RNA in situ hybridization showed the transcriptional<br />

induction <strong>of</strong> two maize defense genes, PRms (Pathogenesis related<br />

protein, maize seeds) and UGT (UDP-glucosyltransferases), in specific seed tissues<br />

in advance <strong>of</strong> visible fungal colonization, suggesting that the fungal virulent<br />

factors or plant defense signaling transduction molecules may be involved. Induction<br />

<strong>of</strong> gene transcription <strong>of</strong>ten appeared first in the tip <strong>of</strong> scutellum, which is<br />

the cotyledon <strong>of</strong> the seed. This observation suggests that the scutellum may be<br />

very important in plant-microbe communication. Even though these two fungi are<br />

characterized as having different trophic lifestyles, they follow a similar pattern <strong>of</strong><br />

colonization, and induce two known defense genes in maize seeds. Perhaps structural<br />

or chemical characteristics <strong>of</strong> seeds favor a common path <strong>of</strong> colonization, or<br />

the two fungi share a common toolbox for the infection <strong>of</strong> maize seeds.<br />

Sikhakolli, Usha 1 , Nicholas Harrison 1 , Nina Lehr 2 , Zheng Wang 2 , Ning Li 2 ,<br />

Frances Trail 1* , and Jeffrey P Townsend 2 . 1 Department <strong>of</strong> Plant Biology,<br />

Michigan State University, East Lansing, MI, 2 Department <strong>of</strong> Ecology and Evolutionary<br />

Biology, Yale Univeristy, New Haven CT. Comparative transcriptomics<br />

identifies new genes important to perithecium development and function<br />

in Fusarium<br />

In recent years, a plethora <strong>of</strong> genomic sequences have been released for<br />

fungal species, accompanied by functional predictions for genes based on protein<br />

sequence comparisons. However, identification <strong>of</strong> genes involved in particular<br />

processes has been extremely slow, and new methodologies for identifying genes<br />

involved in a particular process have not kept pace with the exponential increase<br />

in genome sequence availability. We have performed transcriptional pr<strong>of</strong>iling <strong>of</strong><br />

five species <strong>of</strong> Neurospora and Fusarium during six stages <strong>of</strong> perithecium development.<br />

We estimated the ancestral transcriptional shifts during the developmental<br />

process among the species and identified genes whose transcription had substantially<br />

and significantly shifted during the evolutionary process. We examined<br />

genes whose expression greatly increased in Fusarium graminearum perithecium<br />

development compared to Fusarium verticilloides or compared to Neurospora<br />

spp. Functional studies through gene disruption resulted in substantial changes in<br />

Fusarium graminearum perithecium morphology in the mutants <strong>of</strong> many <strong>of</strong> these<br />

genes. These genes were not previously identified as candidates for function in<br />

perithecium development, illustrating the utility <strong>of</strong> this method for identification<br />

<strong>of</strong> genes associated with specific functional processes.<br />

Continued on following page


Slot, Jason C, Matthew A Campbell, Han Zhang, John G Gibbons, and Antonis<br />

Rokas. Department <strong>of</strong> Biological Sciences, Vanderbilt University, Nashville, TN.<br />

Fungal gene clusters, transfer, tinkering and death<br />

The association <strong>of</strong> metabolically linked genes in chromosomal gene clusters<br />

has been shown to facilitate accelerated evolution <strong>of</strong> fungal genomes through<br />

whole pathway acquisition by horizontal gene transfer (HGT) and through accelerated<br />

pathway loss. How gene clusters originate and integrate into novel fungal<br />

genomes following HGT is not well understood, however. In order to address<br />

these questions, we investigated the evolution <strong>of</strong> several secondary metabolite<br />

gene clusters with highly restricted distributions among closely related fungal<br />

taxa. Here we present evidence that novel gene clusters can persist along with alternative<br />

assemblages <strong>of</strong> genes in polymorphic loci. These variable loci could explain<br />

the dramatic differences in secondary metabolic diversity between closely<br />

related fungi. Recently acquired clusters may alternatively be scrapped, with some<br />

components being repurposed and the rest left to decay. These results provide<br />

more insight into the importance <strong>of</strong> gene clusters in fungal metabolic diversity and<br />

suggest that the rate <strong>of</strong> gene cluster HGT is likely underestimated.<br />

Smith, Matthew E 1 , Terry W Henkel 2 , H David Clarke 3 , Alex K Fremier 4 , and<br />

Rytas Vilgalys 5 . 1 Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Florida,<br />

Gainesville FL 32611, 2 Department <strong>of</strong> Biological Sciences, Humboldt State University,<br />

Arcata CA 95521, 3 Department <strong>of</strong> Biology, University <strong>of</strong> North Carolina-Asheville,<br />

Asheville, NC 28804, 4 Department <strong>of</strong> Fish and Wildlife Resources,<br />

University <strong>of</strong> Idaho, Moscow ID 83844, 5 Department <strong>of</strong> Biology, Duke University,<br />

Durham NC 27708. Ectomycorrhizal fungi associated with Pakaraimaea<br />

dipterocarpacea, an endemic neotropical dipterocarp<br />

Ectomycorrhizal (ECM) plants and fungi can be well represented in certain<br />

tropical regions. For example, the predominantly paleotropical ECM Dipterocarpaceae<br />

is one <strong>of</strong> the most widespread and speciose tree families in Southeast<br />

Asia. Pakaraimaea dipterocarpacea is one <strong>of</strong> only two species <strong>of</strong> dipterocarp<br />

known from the western hemisphere, and is the only member <strong>of</strong> the monotypic<br />

Dipterocarpaceae subfam. Pakaraimoideae. The species was discovered in<br />

Guyana by Bassett Maguire <strong>of</strong> the New York Botanical Garden in the 1950’s, and<br />

is endemic to the sandstone highlands <strong>of</strong> Guyana and Venezuela in dense stands<br />

<strong>of</strong> multi-stemmed trees. In 2006 Bernard Moyersoen showed that Pakaraimaea<br />

forms ectomycorrhizas and provided evidence for a vicariant distribution relative<br />

to its Old World relatives. Despite the unique phylogenetic position <strong>of</strong> P. dipterocarpacea<br />

within the Dipterocarpaceae and its unique geographical distribution,<br />

little is known about the ECM fungi associated with the species. In 2010 we sampled<br />

ECM fungi on roots <strong>of</strong> P. dipterocarpacea and the leguminous ECM tree Dicymbe<br />

jenmanii where these species co-occur in the Upper Mazaruni River Basin<br />

<strong>of</strong> Guyana. Using molecular methods we documented 52 species <strong>of</strong> ECM fungi<br />

across 11 independent lineages. While many <strong>of</strong> ECM fungi found on<br />

Pakaraimaea roots were conspecific with taxa known from the region as sporocarps,<br />

a number <strong>of</strong> “unknowns” were found across several basidiomycete families<br />

and genera. Due to the phylogenetic distance between host taxa we hypothesized<br />

that P. dipterocarpacea would host ECM fungi not found on D. jenmanii,<br />

and vice versa. While statistical tests suggested that several ECM fungal species<br />

exhibited preference for either P. dipterocarpacea or D. jenmanii, most were generalists.<br />

In addition to these findings, unique ECM fungi collected as sporocarps<br />

in the Pakaraimaea forests will be discussed.<br />

Soares, William RO, Eliane AS Armando, and José C Dianese * . Depto. de Fitopatologia,<br />

Universidade de Brasília, 70910900 Brasília, DF, Brazil. A new<br />

species, and first record <strong>of</strong> Meliola weigeltii var. fraxinifoliae, on anacardiaceous<br />

hosts from the Brazilian Pantanal<br />

A new Meliola species was found on leaves <strong>of</strong> Spondias lutea (Brazilian<br />

plum), a cultivated fruit crop also occurring in wild environments through the<br />

Neotropics. Our specimen was collected in the Pantanal Matogrossense (UB<br />

Mycol. Col. 21456) showing characteristic features related to: mycelial setae up<br />

to 267 _ 5.5-9.5 µm, mostly erect, sometimes curved, with a broad apex when single<br />

containing 2-4 denticles up to 13.5 µm tall or 2-3 furcate with denticles up to<br />

22 µm tall; ascomata 90-162 µm diam., superficial, glabrous, dark brown, globose<br />

to subglobose, spread on the superficial mycelium; ascospores 31-41 _ 9.5-<br />

13 µm, oblong to elliptical with obtuse terminal cells, dark brown when mature,<br />

4-septate, with constrictions at each septum. This specimen morphologically differs<br />

from all known Meliola species, strongly indicating that it belongs in a new<br />

taxon. This is also the fourth record <strong>of</strong> a Meliola species on plants <strong>of</strong> the genus<br />

Spondias. Meliola weigeltii var. fraxinifoliae was found on Astronium fraxinifolium,<br />

which is the first record for the entire Pantanal Region, although it is already<br />

known from the Cerrado and from the Northeastern region <strong>of</strong> Brazil.<br />

Soares, William RO, Maria DM Santos, Érica SC Souza, and José C Dianese * .<br />

Depto. de Fitopatologia, Universidade de Brasília, 70910900 Brasília, DF, Brazil.<br />

A new synnematous hyphomycete parasitic on a Phyllachora species<br />

A new Phyllachora species on leaves <strong>of</strong> an unknown angiosperm was recently<br />

found. However, very seldom the stromata <strong>of</strong> the fungus and its immerse<br />

ascomata were free <strong>of</strong> an absolutely inedited hyphomycete. This new hyperparasite<br />

was studied both using light and scanning electron microscopy, and shown to<br />

invade the Phyllachora ascomata and the surrounding pseudostromata that resulted<br />

from the incorporation <strong>of</strong> the neighboring plant tissue by the host fungus. The<br />

hyphomycete develops into a widespread internal compact hyaline mycelium that<br />

later will grow erumpently through the plant epidermis as well as through the<br />

melanin-impregnated clypeus <strong>of</strong> the fungus. Thus, cylindrical compact synnemata<br />

showing textura intrincata grow to 285-500 µm long 57-70 µm diam., completely<br />

covered, from base to apex, by a multitude <strong>of</strong> closely packed verrucose<br />

polyphialidic conidiogenous cells (1-2 µm diam.). The conidia [(4)-13-18 x 1-3<br />

µm)] formed are hyaline, mostly 1-septate, elliptical to fusiform. Given the absolutely<br />

inedited morphology <strong>of</strong> the hyphomycete, it becomes clear that it has to<br />

be assigned to a new anamorphic genus.<br />

Song, Zewei and Jonathan Schilling. Department <strong>of</strong> Bioproducts and Biosystems<br />

Engineering, University <strong>of</strong> Minnesota, St. Paul, MN 55108. Initial population<br />

abundance shaped the competition between two wood-degrading fungi<br />

Competition <strong>of</strong> wood-degrading fungi determines the structure <strong>of</strong> their<br />

community and the resulting ecological processes. During wood decomposition,<br />

brown and white rot fungi utilize different forms <strong>of</strong> carbon while competing for<br />

the same space. In this research, we studied the effect <strong>of</strong> initial population abundance<br />

on the competition outcome and residue properties <strong>of</strong> two wood-degrading<br />

fungi on four wood types. A microcosm system was used to allow brown and<br />

white rot fungi invade and compete on wood substrates (oak, birch, pine and<br />

spruce). The initial population size was controlled by the area <strong>of</strong> agar inoculated<br />

into the microcosm. A brown rot fungus (Gloeophyllum trabeum) was inoculated<br />

in larger amounts (646mm 2 ) than a white rot fungus (Irpex lacteus, 77mm 2 ), simulating<br />

a stronger brown rot population due to earlier colonization. Decay residues<br />

were harvested after 3 and 8 weeks. A previous experiment showed that when<br />

both fungi were inoculated at the same size, the white rot fungus always dominated<br />

the system. The wood residue after 3 weeks showed distinctive white rot<br />

properties, with low solubility (~20%) and high pH (~4). However, after 8 weeks,<br />

a proportion <strong>of</strong> wood residues showed distinctive brown rot properties, which<br />

were high solubility (~60%) and low pH (~2). Quantitative PCR also showed that<br />

at 3 weeks the white rot fungus was dominant in wood. But after 8 weeks, some<br />

<strong>of</strong> the residues showed much higher percentage <strong>of</strong> the brown rot fungus. It is possible<br />

that the brown rot fungus was unsuccessful in invading wood at an early<br />

stage <strong>of</strong> competition, but was able to survive in feeder strips and outcompeted the<br />

white rot fungus after 8 weeks. Our study indicated that initial population abundance<br />

has a pr<strong>of</strong>ound effect on the competition outcome.<br />

Spakowicz, Daniel J, Jeffrey J Shaw, Rahul S Dalal, Gang Fang, Brian F Dunican,<br />

and Scott A Strobel. Dept <strong>of</strong> Molecular Biophysics and Biochemistry, Yale<br />

University, New Haven, CT 06511. Genomic and transcriptomic analysis <strong>of</strong><br />

the novel endophytic isolate E5202H for the identification <strong>of</strong> natural product<br />

pathways<br />

Endophytes are a diverse niche <strong>of</strong> fungi that produce potentially useful<br />

secondary metabolites. Isolate E5202H is an endophyte that was isolated from the<br />

plant Guazuma ulmifolia (Sterculiaceae) in the Cerro Blanco dry forest <strong>of</strong><br />

Ecuador in 2009. This isolate is likely a new species in the genus Massarina, with<br />

86% similarity in its internal transcribed spacer rDNA sequence to the closest<br />

named organism in GenBank by Smith-Waterman alignment. In addition,<br />

E5202H produces a novel molecule with anti-oomycete activity, shown by mass<br />

spectrometry and NMR to be (2E,4E,6E)-6-methylocta-2,4,6-trienoic acid. Also,<br />

it produces a volatile hydrocarbon with potential bi<strong>of</strong>uel properties, shown by<br />

mass spectrometry and NMR to be (3E,5E,7E)-nona-1,3,5,7-tetraene. Metabolic<br />

labeling studies suggest that (3E,5E,7E)-nona-1,3,5,7-tetraene is produced by a<br />

polyketide synthase-like mechanism. Genomic and transcriptomic analyses <strong>of</strong><br />

E5202H are being pursued to identify the genes involved in the production <strong>of</strong><br />

these molecules. The E5202H genome has been assembled to 20 scaffolds with<br />

an estimated size <strong>of</strong> 52 MB. We pursued transcriptome sequencing from five different<br />

conditions where different volatile metabolites were produced, as determined<br />

by GCMS. Correlating the transcription with metabolite production and<br />

isotopic labeling data allows us to generate hypotheses for the genes responsible<br />

for the production <strong>of</strong> these products. The functions <strong>of</strong> candidate genes will be determined<br />

by heterologous expression studies.<br />

Spiegel, Frederick W. Department <strong>of</strong> Biological Sciences, University <strong>of</strong><br />

Arkansas, Fayetteville, AR 72701, USA. Nucletmycea, Holozoa, Opisthokonta,<br />

and the problem with the “Entagled Bank”<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 45


In the famous last paragraph <strong>of</strong> the Origin, Darwin states, “There is<br />

grandeur in this view <strong>of</strong> life...that...from so simple a beginning endless forms most<br />

beautiful and most wonderful have been, and are being, evolved.” From this, in<br />

part, comes the hard to shake idea that the progenitors <strong>of</strong> lineages were both primitive<br />

and simple. By definition, the characters <strong>of</strong> the ancestor <strong>of</strong> a lineage are primitive.<br />

However, many aspects <strong>of</strong> an ancestor, even a deep ancestor, were very<br />

complex, not simple at all. The last common ancestor <strong>of</strong> living eukaryotes was a<br />

sexual organism that had to have had a flagellate stage in its life cycle. These characters<br />

had to have been true <strong>of</strong> the first opisthokont, and it was certainly more<br />

complex than many members <strong>of</strong> the extant opisthokonts. Rather than worry about<br />

whole organisms, perhaps it is better to compare sets <strong>of</strong> characters that we can hypothesize<br />

result from share ancestry. For instance, multicellularity has arisen numerous<br />

times in eukaryotes. In opisthokonts, multicellularity that includes cell migration<br />

has arisen at least three times, at least once in Holozoa (Metazoa), and at<br />

least twice in Nucletmycea (the sorocarpic Fonticula and the dikaryomycetes).<br />

Do the seeds for using cell migration in multicellular development come from<br />

characters <strong>of</strong> the last common ancestor <strong>of</strong> the extant opisthokonts? We do not yet<br />

know, but if we learn to “think outside the box” as comparative biologists, it will<br />

become easier to recognize those characters that need to be compared and to develop<br />

a set <strong>of</strong> model organisms and that allow hypotheses to be tested.<br />

Stajich, Jason E 1 , David A Hewitt 2 , and Gregory Jedd 3 . 1 Plant Pathology and<br />

Microbiology, University <strong>of</strong> California, Riverside, CA USA, 2 Department <strong>of</strong><br />

Botany,<br />

3<br />

Academy <strong>of</strong> Natural Sciences <strong>of</strong> Philadelphia, Philadelphia, PA,<br />

Temasek Life Sciences Laboratory and Department <strong>of</strong> Biological Sciences, The<br />

National University <strong>of</strong> Singapore, Singapore. Insights into independent origins<br />

<strong>of</strong> multicellularity in Neolecta from comparative genomics<br />

The evolution <strong>of</strong> multicellular tissues in fungi, from the most fundamental<br />

aspects <strong>of</strong> hyphal growth to development <strong>of</strong> complex fruiting bodies, required<br />

a host <strong>of</strong> changes in cellular biology <strong>of</strong> fungi. Genome sequence comparisons provide<br />

a means to study the molecular evolutionary changes across species through<br />

the inventories <strong>of</strong> gene content and genome organization. Phylogenomic comparison<br />

<strong>of</strong> genes in the filamentous Pezizomycotina fungi reveals a collection <strong>of</strong> lineage<br />

specific genes including some known to be important for the filamentous<br />

lifestyle such as the Hex proteins <strong>of</strong> Woronin body organelles, which are necessary<br />

for septal pore plugging after hyphal wounding. Neolecta, a Taphrinamycotina<br />

fungus, is phylogenetically distinct from the Pezizomycotina and may represent<br />

an independent lineage where hyphal growth and complex fruiting bodies<br />

evolved. By sequencing and comparing the genome <strong>of</strong> Neolecta irregularis to the<br />

rest <strong>of</strong> the Ascomycete fungi we have found further evidence that filamentousspecific<br />

genes in Pezizomycotina remain specific to the clade. Microscopy reveals<br />

Wornin-body like structures in the hyphae <strong>of</strong> Neolecta, yet Woronin body genes<br />

hex, wsc, leashin and spa9 are not found, suggesting independent origins for this<br />

adaptation. The NADPH oxidases appear to be expressed in all multicellular fruiting<br />

bodies, and one <strong>of</strong> these is encoded in the Neolecta genome, suggesting shared<br />

traits that might have been present in a complex Dikarya ancestor. These and ongoing<br />

analyses will reveal further patterns <strong>of</strong> genome evolution that link cell biology<br />

changes to evolutionary transitions between unicellular and multicellular<br />

forms in fungal history.<br />

Sthultz, Christopher M 1 , Linda TA Van Diepen 2 , Serita D Frey 2 , and Anne<br />

Pringle 1 . 1 Organismic & Evolutionary Biology Harvard University 16 Divinity<br />

Avenue Cambridge, MA 02138, 2 Department <strong>of</strong> Natural Resources and the Environment<br />

University <strong>of</strong> New Hampshire 56 College Road James Hall Durham,<br />

NH 03801. Influences <strong>of</strong> nitrogen deposition and soil warming on saprophytic<br />

fungal community structure, fungal growth, and litter decomposition<br />

Impacts <strong>of</strong> global change on microbial communities remain poorly understood.<br />

However, because microbes are important drivers <strong>of</strong> biogeochemical<br />

cycles, anthropogenically mediated changes may have strong influences on<br />

ecosystem function. Saprophytic fungi are primarily responsible for decomposition<br />

in temperate forest systems, yet relatively little is known about whether global<br />

changes, including nitrogen deposition and soil warming, will alter community<br />

assembly processes. Here we focus on results from the first year <strong>of</strong> a multi-year<br />

litterbag/decomposition experiment using long-term experimental N addition<br />

(control, low and high N treatments) and soil warming plots (control and elevated<br />

5 o C treatments) in a northeastern hardwood forest. We examined saprophytic<br />

fungal community structure, fungal growth, and decomposition from 95 litter<br />

bags harvested after 1 year in the field. Using metagenomics, culturing, and laboratory<br />

microcosms we tested the influence <strong>of</strong> nitrogen deposition and soil warming<br />

on fungal communities and ecosystem function. We document four major patterns:<br />

1) Fungal communities in elevated nitrogen deposition and soil warming<br />

plots were different from control plots. 2) Decomposition varied between treatment<br />

and control plots in both experiments. 3) In the nitrogen experiment fungal<br />

community composition and decomposition were influenced not only by treatment<br />

but also by the source <strong>of</strong> the litter in the bags (from control, low, or high N<br />

plots). There was a significant interaction effect on the community composition.<br />

46 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

4) There were differences in growth rates <strong>of</strong> the same fungal species cultured from<br />

each <strong>of</strong> the three nitrogen treatments. Our results increase understanding <strong>of</strong> the<br />

ecology and evolution <strong>of</strong> saprophytic fungi in a global change context, and add to<br />

what is known about the biodiversity <strong>of</strong> decomposer fungi. In addition, our results<br />

suggest global change may influence the fungi found in the environment, but also<br />

the evolution <strong>of</strong> some fungal species.<br />

Sweeney, Katarina, Michael Freitag, and Jeffrey Stone. Oregon State University,<br />

Department <strong>of</strong> Botany and Plant Pathology, 2082 Cordley Hall, Corvallis, OR<br />

97331-2902. Estimating genetic diversity <strong>of</strong> Cronartium ribicola populations<br />

in Oregon by RAD marker genotyping<br />

The pathogen, Cronartium ribicola, was introduced to North <strong>America</strong> in<br />

1910 and found in SW British Columbia and NW Washington by 1922. Evidence<br />

<strong>of</strong> evolutionary change in North <strong>America</strong>n C. ribicola since its introduction is apparent<br />

and spread <strong>of</strong> two virulent races pathogenic to WPBR-resistant genotypes<br />

<strong>of</strong> Pinus monticola and P. lambertiana have been observed. The C. ribicola vcr1<br />

race is virulent to P. lambertiana carrying a major gene (Cr1) that confers resistance<br />

to “wild-type” C. ribicola and was first found in California in 1978. The vcr2<br />

race is virulent to P. monticola carrying Cr2 that also confers resistance to “wildtype”<br />

C. ribicola, and was found in Oregon in the early 1970s. Neither vcr1 or<br />

vcr2 races are virulent to the resistant genotypes <strong>of</strong> the other species. Few studies<br />

have been completed to estimate the genetic diversity <strong>of</strong> natural populations <strong>of</strong> C.<br />

ribicola in western North <strong>America</strong>, but accurate characterization <strong>of</strong> host phenotypes<br />

in resistance breeding programs requires more detailed understanding <strong>of</strong> the<br />

diversity <strong>of</strong> C. ribicola populations. Resistance screening is carried out with wild<br />

populations known to be predominantly vcr1, vcr2, or “wild-type” C. ribicola. In<br />

order to gain a more precise understanding <strong>of</strong> the molecular basis <strong>of</strong> natural variation<br />

in wild-type strains <strong>of</strong> C. ribicola in Oregon, we are employing restriction<br />

site associated DNA (RAD) marker genotyping <strong>of</strong> C. ribicola aeciospores. RAD<br />

sequencing generates short sequence reads adjacent to restriction enzyme sites<br />

that serve to identify restriction polymorphisms between samples or can be used<br />

to identify the presence <strong>of</strong> SNPs. We are using aeciospore samples from multiple<br />

locations in Oregon. RAD mapping will allow us to relate the variation seen in<br />

pine host reactions to infection by C. ribicola with informative SNPs in the<br />

pathogen.<br />

Sylvain, Iman A and Timothy Y James. University <strong>of</strong> Michigan, Department <strong>of</strong><br />

Ecology and Evolutionary Biology. 830 North University Ann Arbor, MI 48109.<br />

Global population genetic structure <strong>of</strong> Aspergillus niger and Eurotium<br />

rubrum in green c<strong>of</strong>fee beans<br />

Whether small species are ubiquitous and cosmopolitan remains a significant<br />

question in mycological research. Whether “everything is everywhere,” and<br />

how habitat properties or historical contingency influence the distribution <strong>of</strong> fungi<br />

has yet to be fully understood. Determining species ranges for fungi can be challenged<br />

by reliance on morphological characters for the identification <strong>of</strong> species,<br />

and enlisting molecular techniques has shown that fungi may have varied ranges<br />

that scale from largely global to narrowly endemic. Fungi associated with highly<br />

mobile agricultural products may have global distributions that mirror the trade <strong>of</strong><br />

their plant hosts, or show patterns <strong>of</strong> biogeographic structure influenced by the<br />

ecology <strong>of</strong> distinct agroecosystems. Trans-global gene flow in fungi may be facilitated<br />

by the movement <strong>of</strong> agricultural commodities, resulting in panmictic<br />

populations, and the observation that fungi show no correspondence between genetic<br />

and geographic ranges. We are developing global c<strong>of</strong>fee agriculture as a<br />

model system to understand how the movement <strong>of</strong> food products influences the<br />

dispersal <strong>of</strong> fungi. We have used a multi-locus sequencing technique to analyze<br />

the population structure <strong>of</strong> two mycotoxigenic Aspergillus species, A. niger and<br />

Eurotium rubrum, cultured from green c<strong>of</strong>fee beans. C<strong>of</strong>fee beans were sourced<br />

from eight countries: Kenya, Ethiopia, Mexico, Costa Rica, Guatemala, Papa<br />

New Guinea, Bali, and Sumatra. The c<strong>of</strong>fee was either produced as USDA-certified<br />

organic, or by conventional agricultural methods, and either wet or dryprocessed.<br />

Phylogenetic trees constructed from rRNA ITS sequences suggest that<br />

multiple distinct, possibly cryptic species are present within the sampled A. niger<br />

and E. rubrum clades. Beta-tubulin and calmodulin sequences show no geographic<br />

structure in these species, suggesting that Aspergillus species associated<br />

with c<strong>of</strong>fee production may truly be everywhere.<br />

Taerum, Stephen 1 , Z Wilhelm de Beer 2 , Tuan A Duong 1 , Min Lu 3 , Nancy<br />

Gilette 4 , Jianghua Sun 3 , and Michael J Wingfield 1 . 1 Department <strong>of</strong> Genetics,<br />

Forestry and Agricultural Biotechnology Institute (FABI), University <strong>of</strong> Pretoria,<br />

Pretoria 002, South Africa, 2 Department <strong>of</strong> Microbiology and Plant Pathology,<br />

Forestry and Agricultural Biotechnology Institute (FABI), University <strong>of</strong> Pretoria,<br />

Pretoria 002, South Africa, 3 State Key Laboratory <strong>of</strong> Integrated Management <strong>of</strong><br />

Pest Insects and Rodents, Institute <strong>of</strong> Zoology, Chinese Academy <strong>of</strong> Sciences,<br />

Beijing 100101, P. R. China, 4 Ecosystem Function and Health, PSW Research<br />

Station, 800 Buchanan Street, Albany, CA 94710. Fungal symbionts suggest an<br />

Continued on following page


alternative origin for the red turpentine beetle (Dendroctonus valens) invasion<br />

in China<br />

Species invasions are among the greatest threats to biodiversity, ecosystem<br />

stability, and the global economy. Invasive species frequently co-invade with<br />

numerous symbionts, some <strong>of</strong> which are themselves problematic in their new environments.<br />

Interestingly, the symbionts associated with an invading species can<br />

be useful in discerning the possible origins <strong>of</strong> invasive pests. We used phylogenetic<br />

analyses to determine and compare the composition <strong>of</strong> fungal communities<br />

associated with a forest pest, the red turpentine beetle (Dendroctonus valens), in<br />

its native [eastern (ENA) and western (WNA) North <strong>America</strong>] and invaded<br />

(China) environments. In total, 28 different species <strong>of</strong> Ophiostomatalean fungi<br />

were isolated in the three regions: 14 spp. in China, 13 in ENA and 11 in WNA.<br />

Five fungal species were shared between RTB populations in ENA and WNA.<br />

Two species (O. ips and O. floccosum) occurred in both China and WNA, but<br />

both were present in very low numbers and have a global distribution. Three<br />

species were present in ENA and China, including O. abietinum, also with a global<br />

distribution, L. koreanum, thus far only reported from East Asia, and L. procerum.<br />

The first two species respectively constituted 1% and 6% <strong>of</strong> the ophiostomatalean<br />

isolates obtained in both regions, while L. procerum was the dominant<br />

fungus in both regions, constituting 50% <strong>of</strong> the isolates in ENA and 56% in China.<br />

This fungus is pathogenic to pine trees in China, but not in the USA, and was conspicuously<br />

absent from RTB in WNA. These findings suggest that the RTB may<br />

have been introduced into China from ENA, contrary to previous reports based on<br />

population studies <strong>of</strong> the beetle, suggesting that RTB was introduced into China<br />

from WNA.<br />

Talbot, Jenny, Lotus L<strong>of</strong>gren, and Kabir Peay. University <strong>of</strong> Minnesota Plant<br />

Pathology 1991 Upper Buford Circle 495 Borlaug Hall St. Paul, MN 55108-<br />

6030A. Degradative enzyme production: a function <strong>of</strong> phylogenetic or niche<br />

relatedness?<br />

Fungi produce an array <strong>of</strong> extracellular enzymes that assist in the degradation<br />

<strong>of</strong> organic materials. Fungi occupying different ecological niches are expected<br />

to have different enzymatic capabilities but the extent to which these capabilities<br />

are restricted by genetic lineage is unclear. We are testing the hypothesis<br />

that fungi occupying similar ecological niches have similar enzyme pr<strong>of</strong>iles but<br />

that this capability is restricted by genetic lineage. To test this hypothesis, we are<br />

using 18 sets <strong>of</strong> closely related fungi across the Basidiomycota in a laboratorybased<br />

microcosm study to directly quantify the production <strong>of</strong> degradative enzymes<br />

among functionally and phylogenetically distinct taxa. Each set consists <strong>of</strong><br />

‘three pairs’ <strong>of</strong> fungi occupying distinct ecological niches: a mycorrhizal fungus,<br />

a white rot fungus, and a brown rot fungus, for a total <strong>of</strong> 54 taxa. We are measuring<br />

the production <strong>of</strong> cellulases, hemicellulases, proteases, and oxidases by each<br />

taxa as they colonize litter and soil in axenic culture. We will then correlate genetic<br />

distance to functional distance among taxa and test the effect <strong>of</strong> ecological<br />

niche on enzyme production using ANOVA. ECM fungi form associations with<br />

host plants under which fungal-sequestered nutrients are exchanged for plantfixed<br />

carbon. By contrast, white rot fungi and brown rot fungi acquire carbon<br />

from detritus, with white rot fungi having stronger lignin degrading capabilities<br />

than brown rot. Viewing enzyme production as an adaptive functional trait, we expect<br />

that the enzymatic pr<strong>of</strong>iles <strong>of</strong> different fungal species will vary most strongly<br />

with ecological niche, with a lesser effect <strong>of</strong> phylogenetic distance between lineages.<br />

However, we expect that closely related species will produce a similar suite<br />

<strong>of</strong> enzymes.<br />

Tanney, Joey B and Keith A Seifert. Eastern Cereal and Oilseed Research Centre,<br />

960 Carling Avenue, Ottawa, Ontario, K1A 0C6. The occurrence <strong>of</strong><br />

Phialosimplex in the built environment<br />

House dust samples were collected from 14 countries as part <strong>of</strong> an ongoing<br />

project comparing the indoor mycota detected by dilution-to-extinction culturing<br />

and pyrosequencing. A modified isolation method using microtubes and<br />

low water activity agar media was used to culture slow-growing and xerophilic<br />

fungi that would potentially be missed by classical methods, which favor fastgrowing<br />

mesophilic molds. Strains identified as Phialosimplex comprised approximately<br />

15% <strong>of</strong> identified cultures from house dust samples from Micronesia,<br />

0.8% from Indonesia, and an as yet unquantified percentage <strong>of</strong> strains from<br />

Thailand. Phialosimplex is a recently described anamorphic genus within the Trichocomaceae.<br />

Morphologically, it is characterized by narrow phialides borne on<br />

vegetative hyphae or on short, unbranched conidiophores, with aseptate subglobose<br />

to ovoid conidia in long chains. There are three known species, P. caninus,<br />

P. chlamydosporus, both associated with disseminated mycoses in dogs, and P.<br />

sclerotialis, originally isolated from hay. In a recently published study, the latter<br />

species comprised 37% <strong>of</strong> fungi identified in a poultry facility using pyrosequencing.<br />

Preliminary ITS data suggest the presence <strong>of</strong> at least one novel<br />

Phialosimplex species among the house dust isolates. Our ITS sequences <strong>of</strong><br />

Phialosimplex will be used to identify OTUs in the 454 pyrosequencing data generated<br />

from the same dust samples, allowing us to compare the sensitivity <strong>of</strong> the<br />

microbiological and molecular methods for detecting species <strong>of</strong> this genus. The<br />

frequent occurrence in house dust <strong>of</strong> Phialosimplex, a genus previously considered<br />

as an obscure agent <strong>of</strong> canine disease, should encourage further investigation<br />

into its significance in the built environment.<br />

Taylor, Andy F S 1,2 , Greg W Douhan 3 , Allan E Hills 4 , Giampaolo Simonini 5 ,<br />

Manfred Binder 6 , and Ursula Eberhardt 7 . 1 The James Hutton Institute,<br />

Craigiebuckler, Aberdeen, AB15 8QH, Scotland, UK, 2 Institute <strong>of</strong> Biological and<br />

Environmental Sciences, University <strong>of</strong> Aberdeen, Cruickshank Building, St<br />

Machar Drive, Aberdeen, AB24 3UU, UK, 3 Department <strong>of</strong> Plant Pathology and<br />

Microbiology, University <strong>of</strong> California, Riverside, CA 92521, USA, 4 ‘Megera’,<br />

Acremead Road, Wheatley, Oxon, OX33 1NZ, England, UK, 5 Via Bellaria 8,<br />

42100 Reggio Emilia, Italy, 6 Clark University, Biology Department, Lasry Biosciences<br />

7<br />

Center, 950 Main Street Worcester, Massachusetts 01610-1477, U.S.A.,<br />

CBS-KNAW Fungal Biodiversity Centre, Centraalbureau voor Schimmelcultures,<br />

P. O. Box 85167, NL-3508 AD UTRECHT, The Netherlands. A molecular<br />

analysis <strong>of</strong> European and North <strong>America</strong>n taxa within the Xerocomus<br />

chrysenteron complex and the description <strong>of</strong> X. redeuilhii Taylor, Eberhardt<br />

& Simonini sp. nov.<br />

The taxonomy within the genus Xerocomus has not been fully resolved<br />

and remains controversial. For example, in North <strong>America</strong>, Smith and Thiers collapsed<br />

all species within the genus Xerocomus into the genus Boletus in the<br />

1970’s. Nevertheless, Xerocomus is widely recognized as a distinct genus in Europe<br />

and phylogenetic analysis based on LSU-rDNA supports that Xerocomus is<br />

a distinct genus. Phylogenetic analyses have also revealed that the type species for<br />

the genus, Xerocomus subtomentosus, is distinct from other Xerocomus species<br />

that have been studied at the molecular level, which suggests that additional genera<br />

could be split <strong>of</strong>f. In this paper we examine the taxa in the Xerocomus chrysenteron<br />

complex in Europe and North <strong>America</strong> using rDNA-ITS sequence data.<br />

We describe Xerocomus redeuilhii A.F.S. Taylor, Eberhardt & Simonini sp. nov.,<br />

a widespread species from southern Europe that was formerly considered to be the<br />

North <strong>America</strong>n species X. dryophilus (Theirs) Singer. Xerocomus fennicus (Harmaja)<br />

Ladurner & Simonini described from Finland is shown to be synonymous<br />

with the North <strong>America</strong>n taxon Xerocomus intermedius (Smith & Thiers) Heinemann,<br />

Rammeloo &. Rullier and X. marekii Šutara & Skála is shown to be a red<br />

form <strong>of</strong> X. porosporus Imler. Phylogenetic analysis <strong>of</strong> the complex demonstrated<br />

that few taxa are shared between North <strong>America</strong> and Europe, with a number <strong>of</strong><br />

lineages within the complex requiring taxonomic re-evaluation in North <strong>America</strong>.<br />

Taylor, Lee 1 , Jack W McFarland 2 , Teresa N Hollingsworth 3 , and Roger W<br />

Ruess 1 . 1 Institute <strong>of</strong> Arctic Biology, 311 Irving I Building, University <strong>of</strong> Alaska<br />

Fairbanks, Fairbanks, AK 99775, 2 U.S. Geological Survey, 345 Middlefield<br />

Road, M.S. 962, Menlo Park, CA 94025, 3 Boreal Ecology Cooperative Research<br />

Unit, PNW Research Station, USDA Forest Service, Fairbanks, AK 99775. Saturated<br />

molecular census <strong>of</strong> soil fungi in boreal Alaska suggests 6 million<br />

species <strong>of</strong> fungi on earth<br />

There are 98,000 described fungal species and the true number <strong>of</strong> fungal<br />

species is estimated to range from 0.6M based on a comparison <strong>of</strong> animal species<br />

to 1.5 M based on the number <strong>of</strong> pathogenic fungi associated with plants. Because<br />

systematics in fungi and other microbes is vastly different from that in animals,<br />

because most fungi are not plant pathogens, and because molecular methods were<br />

not used, existing estimates are likely to be very wrong. We used an environmental,<br />

molecular approach to census plant and fungal diversity in a simple<br />

ecosystem, with particular concern for adequacy <strong>of</strong> sampling and accuracy <strong>of</strong><br />

species assessment. Our estimate, based on fungi found in soil associated with<br />

plants at a ratio <strong>of</strong> 17 fungi to 1 plant, is 6M fungal species, itself a conservative<br />

estimate. Our result indicates fungal diversity has been greatly underestimated.<br />

Our data <strong>of</strong>fer an explanation for this underestimate; niche partitioning on a very<br />

fine scale. Furthermore, we find that species in the same genus and broad functional<br />

guild <strong>of</strong>ten occupy different niches. With the ongoing reorganization <strong>of</strong> biodiversity<br />

on Earth, it is critical to accelerate our documentation <strong>of</strong> the distribution<br />

and function <strong>of</strong> the likely 98% <strong>of</strong> fungal species that are currently undescribed.<br />

Taylor, Steven J 1 , Andrew N Miller 1* , Anthony C Yannarell 2 , Joseph F Merritt<br />

1 , Nohra Mateus-Pinilla 1 , Edward J Heske 1 , Vincent P Hustad 3 , H M Lin 2 ,<br />

Joseph A Kath 4 , and Rod D McClanahan 5 . 1 Illinois Natural History Survey, University<br />

<strong>of</strong> Illinois, 1816 South Oak St., Champaign, IL 61820, 2 Dept. <strong>of</strong> Natural<br />

Resources and Environmental Science, University <strong>of</strong> Illinois, 1102 South Goodwin<br />

Ave., Urbana, IL 61801, 3 Dept. <strong>of</strong> Plant Biology, University <strong>of</strong> Illinois, 505<br />

South Goodwin Ave., Urbana, IL 61801, 4 Illinois Department <strong>of</strong> Natural Resources,<br />

One Natural Resources Way, Springfield, IL 62702, 5 US Forest Service,<br />

Shawnee National Forest - Hidden Springs RD, 602 North First St., Vienna, IL<br />

62995. White-nose syndrome and Illinois bat hibernacula<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 47


We are monitoring the invasion <strong>of</strong> Geomyces destructans in bat hibernacula<br />

in Illinois. Our team includes three mammalogists, a mycologist, a microbial<br />

ecologist, a wildlife veterinary epidemiologist, and a cave biologist, and is assisted<br />

by several resource managers. We use molecular and culture-based<br />

approaches to evaluate dead and live-caught bats and cave and mine substrates for<br />

the presence <strong>of</strong> G. destructans, and describe the microbial and fungal communities<br />

<strong>of</strong> sampled animals and caves. Beginning in winter 2012, we will visit about<br />

eight hibernacula per year for three years, and sample active bats during summer.<br />

We are collecting swab and wing-punch samples from asymptomatic and symptomatic<br />

bats; soil, air, and various other substrate samples from hibernacula; and<br />

temperature, humidity, and light data to characterize cave environments. Our<br />

study will provide data on the occurrence and distribution <strong>of</strong> G. destructans in hibernacula<br />

on the leading edge <strong>of</strong> the spread <strong>of</strong> white-nose syndrome, and the microbial<br />

ecosystems in which it becomes established. Circumstances permitting<br />

(i.e., timing and extent <strong>of</strong> the invasion), we hope to better understand potential<br />

competing or synergistic interactions between G. destructans and microbial communities<br />

on bats and in caves that influence the establishment <strong>of</strong> G. destructans.<br />

Thiers, Barbara M, Deb Paul, Cathy Bester, and Jason Grabon. iDigBio,<br />

FLMNH, Dickinson Hall, Gainesville, FL 32611. Integrated Digitized Biocollections<br />

(iDigBio): A national resource for digitization<br />

iDigBio, Integrated Digitized Biocollections, is the National Resource<br />

funded by the National Science Foundation for Advancing Digitization <strong>of</strong> Biological<br />

Collections (ADBC). Through iDigBio, data and images for millions <strong>of</strong> biological<br />

specimens are being curated, connected and made available in electronic<br />

format for the biological research community, government agencies, students,<br />

educators, and the general public. The mission <strong>of</strong> iDigBio is to develop a national<br />

infrastructure that supports the vision <strong>of</strong> ADBC by overseeing implementation<br />

<strong>of</strong> standards and best practices for digitization; building and deploying a customized<br />

cloud computing environment for collections; recruiting and training personnel,<br />

including underserved groups; engaging the research community, collections<br />

community, citizen scientists, and the public through education and outreach<br />

activities; and planning for long-term sustainability <strong>of</strong> the national digitization effort.<br />

In addition to the iDigBio central digitization HUB, there are partner institutions<br />

referred to as “Thematic Collection Networks” (TCNs) that consists <strong>of</strong> networks<br />

<strong>of</strong> institutions with a strategy for digitizing information that addresses a<br />

particular research theme. The current TCNs include InvertNet - An Integrative<br />

Platform for Research on Environmental Change, Species Discovery and Identification;<br />

Plants, Herbivores and Parasitoids: A Model System for the Study <strong>of</strong> Tri-<br />

Trophic Associations; and North <strong>America</strong>n Lichens and Bryophytes: Sensitive Indicators<br />

<strong>of</strong> Environmental Quality and Change. Additional TCNs will be added<br />

yearly, with four new TCNs to be announced in April 2012. Through the iDigBio<br />

HUB cyberinfrastructure, compilation and the inter-linking <strong>of</strong> data from the<br />

TCNs and existing collaborative databases, there will be opportunities to address<br />

research questions and education interests regarding biodiversity, climate change,<br />

species invasions, natural disasters, and the spread <strong>of</strong> pests and diseases. New<br />

TCNs will be funded in succeeding years based on solicitations from NSF. The<br />

iDigBio HUB is based at the University <strong>of</strong> Florida (UF), in partnership with Florida<br />

State University (FSU).<br />

Thomas, Elizabeth, Nick Taylor, Pio Moises Figueroa-Contreras, Zhi Zhang,<br />

Kelly L Ivors, and Marc A Cubeta. Department <strong>of</strong> Plant Pathology, North Carolina<br />

State University, Raleigh, NC 27695. Investigation <strong>of</strong> sclerotial morphogenesis<br />

in the soil fungus Rhizoctonia solani<br />

The soil fungus Rhizoctonia solani (teleomorph=Thanatephorus cucumeris)<br />

is an important pathogen <strong>of</strong> cultivated and native species <strong>of</strong> plants. The<br />

fungus is a competitive saprobe that promotes the decomposition <strong>of</strong> organic matter<br />

and survives in soil in the absence <strong>of</strong> a host plant by forming sclerotia. Although<br />

there has been an accumulating body <strong>of</strong> knowledge generated on sclerotial<br />

forming fungi, little is known about the molecular mechanisms involved in the<br />

formation <strong>of</strong> sclerotia by R. solani. To examine the biochemical and physiological<br />

changes <strong>of</strong> the fungus during sclerotial morphogenesis, an in vitro assay was<br />

developed to induce sclerotia formation in a wild type strain <strong>of</strong> R. solani anastomosis<br />

group 3 (Rhs 1AP) and two protoplast-derived strains with a reduced nuclear<br />

genome (Rhs 1AP-115 and Rhs 1AP-123E) that had lost their ability to produce<br />

monilioid cells and sclerotia on nutrient media and potato tubers. The<br />

addition <strong>of</strong> xylose, arabinose, galactose, mannose, maltose, trehalose, raffinose,<br />

starch, cellulose, xylan oat-spelt, and casein to minimal medium stimulated radial<br />

mycelial growth <strong>of</strong> Rhs1AP but not Rhs 1AP-115 and Rhs 1AP-123E. The supplementation<br />

<strong>of</strong> minimal media with arabinose, mannitol, sorbitol, xylitol, glycerol,<br />

and myo-inositol induced formation <strong>of</strong> sclerotia in strain Rhs 1AP but not in<br />

strains Rhs 1AP-115 and Rhs 1AP-123E. Growth <strong>of</strong> the latter strains in the presence<br />

<strong>of</strong> cAMP or under nutrient deprived conditions also did not result in the formation<br />

<strong>of</strong> sclerotia. Results from the analysis <strong>of</strong> the transcriptome <strong>of</strong> strains Rhs<br />

1AP (sclerotia plus) and Rhs 1AP-123E (sclerotia minus) will be discussed.<br />

48 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Tobias, Terri L 1 , Katrina P Sandona 1 , Antonio D Rosales 1 , Andrea Porras-Alfaro<br />

1 , Robert L Sinsabaugh 2 , and Katharine N Suding 3 . 1 Western Illinois University,<br />

1 University Circle, Macomb IL. 61455, 2 University <strong>of</strong> New Mexico,<br />

167A Castetter Hall, Albuquerque NM. 87131, 3 University <strong>of</strong> California Berkeley,<br />

130 Mulford Hall 3144, Berkeley CA. 94720. In search for a function: Endophytic<br />

fungi in the alpine tundra and their effect on cultivated plants.<br />

Plants in the alpine tundra are anatomically adapted to survive harsh environmental<br />

conditions. They also rely on endophytic and mycorrhizal fungi for<br />

acquisition <strong>of</strong> nutrients and potential protection against pathogens. Microbial interactions<br />

with plants are known to play an important role in maintaining the biodiversity<br />

in the alpine tundra. Besides their importance, very little is known about<br />

fungal endophytic abundance, taxonomic and functional relationships with dominant<br />

plants. The main objective <strong>of</strong> this research is to examine the mechanism <strong>of</strong><br />

transmission, diversity and function <strong>of</strong> seed and root associated fungi in an alpine<br />

tundra site located at the Niwot Long Term Ecological Research Site in Colorado.<br />

Seed endophytes were compared with root and soil communities using sequencing<br />

and culturing techniques. Dominant fungi were also tested in germination experiments.<br />

Dominant plants in alpine tundra harbor a community <strong>of</strong> endophytic<br />

fungi dominated by Ascomycota. A total <strong>of</strong> 33 endophytes were isolated from<br />

surface sterilized seeds. Greenhouse experiments with a dominant root fungal isolate<br />

identified as Philocephala fortinii showed significant positive effects on both<br />

germination and plant growth in two commercially cultivated plants (Zea mays<br />

and Glycine max). The average number <strong>of</strong> roots for soybeans after eleven days<br />

was 8.6 roots per plant as compared to the control <strong>of</strong> only 1.9 roots per plant. Likewise<br />

corn showed similar results with an average <strong>of</strong> 6.8 roots compared to the<br />

controls with 4.8 roots after seven days. Modern agriculture faces the serious issue<br />

<strong>of</strong> feeding a growing population while preserving the environment. This research<br />

will not only help understand the role <strong>of</strong> these fungi in natural ecosystems but has<br />

the potential to provide new insights in plant health and management.<br />

Toome, Merje and M Catherine Aime. Department <strong>of</strong> Plant Pathology and Crop<br />

Physiology, Louisiana State University Agricultural Center, Baton Rouge,<br />

Louisiana 70803, USA. Microbotryomycetes: Higher-level classification and<br />

description <strong>of</strong> a new genus<br />

Microbotryomycetes, i.e., the “anther smuts” are a class <strong>of</strong> diverse Pucciniomycotina<br />

for which higher-level (e.g., order and family) classification has<br />

been largely unresolved. Recently, phylogenetic reevaluation <strong>of</strong> Microbotryomycetes<br />

was prompted by the isolation <strong>of</strong> a previously undescribed ballistosporic<br />

yeast from a neotropical forest in South <strong>America</strong>. The isolate, recovered<br />

from the surface <strong>of</strong> a fern leaf, forms butyrous cream-colored colonies with polar<br />

budding cells on solid media and rosette-like formations in liquid media. Based<br />

on phylogenetic, morphological and physiological analyses a new genus and<br />

species, Blackwellia eburnea nom. prov., are proposed. Phylogenetic analyses<br />

show that B. eburnea is most closely related to Kriegeria eriophori, a sedge parasite,<br />

the aquatic fungus Camptobasidium hydrophilum, and several recently described<br />

anamorphic yeasts that have been isolated from various plant or psychrophilic<br />

environments. Blackwellia eburnea makes an interesting addition to<br />

this clade since it represents the only species known thus far from the neotropics;<br />

all other species are either from temperate or cold environments. This indicates<br />

that additional studies <strong>of</strong> fungal biodiversity from poorly examined habitats like<br />

ice, water and tropical plant surfaces could greatly contribute to the species discovery<br />

and provide valuable information for improving the taxonomy <strong>of</strong> predominantly<br />

anamorphic clades in Microbotryomycetes. A higher-level classification<br />

is proposed to accommodate this and other species that until now have been<br />

incertae sedis within the Microbotryomycetes.<br />

Tournas, V H 1 , J Rivera Calo 2 , and C Sapp 3 . 1 Center for Food Safety and Applied<br />

2<br />

Nutrition/FDA, 5100 Paint Branch Parkway,<br />

3<br />

College Park, MD, USA,<br />

University <strong>of</strong> Arkansas, Fayetteville, AR, USA, Johns Hopkins School <strong>of</strong> Medicine,<br />

Baltimore, MD, USA. Fungal pr<strong>of</strong>iles in various milk thistle botanicals<br />

from the US market<br />

Milk thistle (MT) dietary supplements are widely consumed due to their<br />

possible beneficial effect on liver health. As botanicals, they can be contaminated<br />

with a variety <strong>of</strong> fungi and their secondary metabolites, mycotoxins. This study<br />

was conducted in an effort to determine the mycological levels and pr<strong>of</strong>iles in various<br />

MT botanical supplements (seeds, herb, tea bags, liquid seed extracts, capsules<br />

and s<strong>of</strong>t gels) obtained from U.S. retail. Conventional plating methods were<br />

used for the isolation and enumeration <strong>of</strong> fungi, while conventional microscopy<br />

as well as molecular methods were employed for species-level identification <strong>of</strong><br />

the isolated fungal strains. Results showed that a high percentage (62%) <strong>of</strong> the MT<br />

samples tested were contaminated with fungi. Total yeast and mould (Y&M)<br />

counts ranged between


tentially toxigenic moulds from the Aspergillus flavus and A. niger groups as well<br />

as Eurotium, Penicillium, Fusarium and Alternaria species were isolated from<br />

MT supplements. The predominant moulds were A. flavus, Eurotium spp. (E.<br />

repens, E. amstelodami and E. rubrum), A. tubingensis, A. niger and A. candidus.<br />

To our knowledge, this is the first study reporting on fungal contamination pr<strong>of</strong>iles<br />

<strong>of</strong> MT botanicals.<br />

Tournas, V H 1 , Naomi Santillan 2 , and Nicholas S Niazi 3 . 1 CFSAN/FDA, 5100<br />

Paint Branch Parkway, College Park, MD, USA, 2 HACU/University <strong>of</strong> Texas,<br />

Rio Grande Valley, TX, USA, 3 JIFSAN/University <strong>of</strong> Maryland, College Park,<br />

MD, USA. Viability and efficacy <strong>of</strong> probiotic yeast supplements from U.S. retail<br />

Yeast (Saccharomyces boulardii) probiotic preparations from three U.S.<br />

companies (A, B and C) were tested for viability and efficacy. A total <strong>of</strong> 33 samples<br />

were analyzed for levels <strong>of</strong> live organisms per gram and per capsule using the<br />

dilution plating method. All samples were tested before their expiration dates. Results<br />

showed that overall S. boulardii levels ranged between 2.6 x 10 9 to 4.0 x<br />

10 10 colony forming units per gram (cfu/g) or 1.4 x 10 9 -8.0 x 10 9 cfu/capsule.<br />

Probiotic supplements from company A had S. boulardii levels similar to/or higher<br />

than those on the label, while all samples from company B had significantly<br />

higher counts than the ones claimed on the label. Most samples from company C<br />

contained significantly lower yeast counts than the ones claimed on their respective<br />

labels. Variation in viable yeast levels existed among companies, among lots<br />

from the same company as well as within lots.<br />

Treseder, Kathleen K and Krista L McGuire. Dept <strong>of</strong> Ecology and Evolutionary<br />

Biology, University <strong>of</strong> California Irvine, Irvine CA 92697. The role <strong>of</strong> fungi<br />

in mediating ecosystem responses to global change<br />

We address whether trait-based mechanisms determine which fungal<br />

species respond to global warming, and whether the species that proliferate are<br />

those that will target recalcitrant C compounds, thereby reduce long-term C sequestration.<br />

The phylogenetic breadth <strong>of</strong> each trait should determine the degree to<br />

which the two traits may overlap within species. We hypothesized that larger,<br />

more complex (i.e., more recalcitrant) organic C compounds will be decomposed<br />

by a narrower phylogenetic range <strong>of</strong> taxa than are more labile C compounds, because<br />

the enzymatic machinery required to break down more recalcitrant compounds<br />

may be more complex and should evolve less frequently. Conversely, we<br />

hypothesized that a relatively broad phylogenetic range <strong>of</strong> taxa would respond to<br />

warming, given that multiple physiological changes can confer adaptation. We<br />

tested these predictions in an Alaskan boreal ecosystem. We used DNA sequencing<br />

to identify fungal taxa that were positively or negatively affected by warming.<br />

In addition, we assessed the relative abundance <strong>of</strong> recalcitrant- and labile-C users<br />

in situ by using nucleotide analog labeling <strong>of</strong> DNA. We found that taxa that fungi<br />

that used lignocelluloses and tannin-proteins, which are relatively recalcitrant<br />

compounds, were more phylogenetically narrow than expected by random chance<br />

(net relatedness index or NRI = 0.6 and 1, respectively). Fungal taxa that occupied<br />

warmed plots were neutrally-related to one another, as were those occupying<br />

control plots. The two traits were linked within fungal taxa, with recalcitrant C<br />

users responding positively to warming (P = 0.026). Ultimately, warming may<br />

lead to decreases in long-term C sequestration in the boreal forest.<br />

Tretter, Eric D, Eric M Johnson, Yan Wang, Prasanna Kandel, and Merlin<br />

White. Boise State University, Dept. <strong>of</strong> Biological Sciences, 1910 University<br />

Drive, Boise, ID 83725. Utilizing new genes (MCM7 and TSR1) for the phylogenetic<br />

analysis <strong>of</strong> early-diverging fungi<br />

The Kickxellomycotina is a group <strong>of</strong> early-diverging fungi with highly diverse<br />

morphologies and life histories that were previously part <strong>of</strong> the non-monophyletic<br />

Zygomycota phylum. Phylogenetic studies <strong>of</strong> the Kickxellomycotina<br />

have been limited by our current inability to culture many members <strong>of</strong> the Harpellales<br />

and all members <strong>of</strong> the Asellariales, and the tendency <strong>of</strong> some clades to suffer<br />

from severe long-branch attraction (particularly the Dimargaritales). This reduces<br />

our ability to sample taxa within each order, and makes it difficult to find<br />

congruent gene trees to produce concatenated multi-gene analyses. Traditional<br />

gene loci may also have severe problems with alignment that limit phylogenetic<br />

power (ITS), have problems with paralogs (beta-tubulin), or may prove very difficult<br />

to amplify for samples not taken from culture (RPB1, RPB2). In an attempt<br />

to address these problems and find additional genes to improve our resolution<br />

within this subphylum, we have investigated the use <strong>of</strong> the genes MCM7 and<br />

TSR1, which have been proven to be powerful for phylogenetic inference within<br />

the Dikarya and congruent with trees produced by phylogenomic analysis. We developed<br />

primers suitable for the amplification <strong>of</strong> MCM7 and TSR1, within three<br />

<strong>of</strong> the current orders <strong>of</strong> Kickxellomycotina (Harpellales, Dimargaritales, and<br />

Kickxellales) and four genera that may represent new orders. To determine if<br />

these genes reveal phylogenies congruent with the evolutionary history <strong>of</strong> the organisms,<br />

we compare the TSR1 and MCM7 phylogenies to an analysis based on<br />

ribosomal RNA genes, the most commonly used and accepted gene locus for<br />

these taxa. Finally, we report on the success <strong>of</strong> attempts to amplify species within<br />

other groups <strong>of</strong> early-diverging fungi, and provide suggestions as to which<br />

primers could potentially work with other clades.<br />

Udayanga, Dhanushka 1,2 , Xingzhong Liu 2 , Amy Y Rossman 3 , Ekachai<br />

Chukeatirote 1 , and Kevin D Hyde 1 . 1 Institute <strong>of</strong> Excellence in Fungal Research,<br />

School<br />

2<br />

<strong>of</strong> Science, Mae Fah Luang University, Chiang Rai 57100, Thailand,<br />

State Key Laboratory <strong>of</strong> Mycology, Institute <strong>of</strong> Microbiology, Chinese Academy<br />

<strong>of</strong> Sciences, No 3 1st West Beichen Road, Chaoyang District, Beijing 100101,<br />

P.R. China, 3 Systematic Mycology & Microbiology Laboratory, USDA-ARS,<br />

Beltsville, MD 20705, USA. Taxonomic and molecular phylogenetic studies<br />

<strong>of</strong> the phytopathogenic genus Diaporthe (Phomopsis)<br />

The genus Diaporthe (anamorph Phomopsis) comprises phytopathologically<br />

important micr<strong>of</strong>ungi with a diverse host association and worldwide distribution.<br />

Species recognition <strong>of</strong> Diaporthe has historically been based on morphology,<br />

cultural characteristics and host affiliation. Difficulties in accurate species<br />

identification using morphology have led to the application <strong>of</strong> alternative approaches<br />

to differentiate species; these include virulence and pathogenicity, biochemistry,<br />

metabolites, physiology, antagonism, molecular phylogenetics and<br />

mating. We investigate and evaluate the taxonomy and phylogeny based on specimens<br />

collected from northern Thailand, specimens and cultures obtained from<br />

BPI and various other collections. The complexes <strong>of</strong> species that occur on each <strong>of</strong><br />

the selected hosts and families are also considered. Fresh specimens are identified<br />

by means <strong>of</strong> molecular phylogenetic data based on the ITS rDNA, partial elongation<br />

factor 1-a (EF), B-tubulin (TUB), calmodulin (CAL) coupled with morphology.<br />

Backbone phylogenetic trees are inferred based on type derived sequences<br />

and used to identify the species and define novel taxa and phylogenetic species<br />

recognition <strong>of</strong> Diaporthe is re-evaluated. The older generic name Diaporthe is put<br />

forward and several species combinations and new taxa are proposed. The need<br />

for epitypification <strong>of</strong> prevalent phytopathogens and adoption to a single generic<br />

name is emphasized.<br />

Uehling, Jessie K 1 , Terry W Henkel 1 , Rytas Vilgalys 2 , and Matthew E<br />

Smith 2,3 . 1 Department <strong>of</strong> Biological Sciences, Humboldt State University, Arcata<br />

CA 95521, USA, 2 Department <strong>of</strong> Biology, Duke University, Durham NC<br />

27708, USA, 3 Department <strong>of</strong> Plant Pathology, University <strong>of</strong> Florida, Gainesville<br />

FL 32611, USA. Membranomyces species are common ectomycorrhizal symbionts<br />

across the northern hemisphere<br />

Membranomyces Jülich is an ectomycorrhizal (ECM) resupinate basidiomycete<br />

genus that currently contains two species, M. delectabilis and M.<br />

spurius, known from Europe and Canada. Studies have suggested that Membranomyces<br />

falls within the Clavulinaceae (Cantharellales), which otherwise consists<br />

<strong>of</strong> the largely coralloid genus Clavulina. Some phylogenetic analyses have<br />

resolved Membranomyces as sister to Clavulina, while others suggested that<br />

Membranomyces is nested within Clavulina. With the recent discovery <strong>of</strong> two<br />

tiny, resupinate Clavulina species in Guyana, we hypothesized that some <strong>of</strong> the<br />

“missing” Clavulina species diversity indicated by belowground molecular studies<br />

in Guyana may be composed other similarly cryptic taxa. Given that a number<br />

<strong>of</strong> temperate belowground ECM studies have found Clavulinaceae root tip sequences<br />

that were unidentified at the species level, we suspected that a missing<br />

diversity situation might also apply to Membranomyces. We sequenced ITS<br />

rDNA from herbarium specimens <strong>of</strong> M. delectabilis and M. spurius to provide a<br />

database from which to determine if conspecific sequences were already present<br />

from root or soil samples deposited in GenBank. We found numerous Membranomyces<br />

environmental sequences on GenBank; subsequent phylogenetic analyses<br />

including these and specimen sequences indicated that 1) Membranomyces<br />

species are more widely distributed in the northern hemisphere than previously<br />

thought, forming ectomycorrhizas with a diverse array <strong>of</strong> gymnosperm and angiosperm<br />

host plants, and 2) unidentified Membranomyces species exist. Host<br />

plant genera with Membranomyces associates included Arctostaphylos, Betula,<br />

Carpinus, Fagus, Populus, Quercus, Pinus and Tsuga, with an apparent new distribution<br />

including Europe, the Mediterranean Arc, the Middle East, Japan, New<br />

Zealand, and United States. We discuss the taxonomic, ecological, and biogeographical<br />

implications <strong>of</strong> these findings.<br />

Urbina, Hector 1 , Janet J Luangsa-ard 2 , Sung O Suh 3 , and Meredith Blackwell 1 1<br />

.<br />

Department <strong>of</strong> Biological Sciences, Louisiana State University, Baton Rouge,<br />

Louisiana 70803, USA, 2 Phylogenetics Laboratory, BIOTEC, Thailand Science<br />

Park, 113 Paholyothin Rd., Khlong 1, Khlong Luang, Pathum Thani 12120, Thailand,<br />

3 Mycology and Botany Program, <strong>America</strong>n Type Culture Collection<br />

(ATCC), 10801 University Blvd., Manassas, Virginia 20110, USA. Xylose-fermenting<br />

yeasts from the gut <strong>of</strong> passalid beetles: geographical distribution<br />

Passalid beetles (Passalidae, Polyphaga, Coleoptera) have a worldwide<br />

distribution with approximately 960 species classified in five subfamilies. Our<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 49


previous studies <strong>of</strong> gut symbionts <strong>of</strong> Odontotaenius disjunctus, a common passalid<br />

beetle in the eastern USA, showed that several xylose-fermenting (X-F)<br />

yeasts are almost always associated with the beetle species. However, similar information<br />

on yeast associates <strong>of</strong> other passalids remains limited. In order to understand<br />

more about the associations between these beetles and X-F yeasts, a variety<br />

<strong>of</strong> passalids were collected from Guatemala and Thailand. Over 2500 yeast<br />

strains were isolated from about 200 adult beetles belonging to 32 passalid species<br />

collected in the two countries. Multi-locus phylogenetic analyses and biochemical<br />

characterizations indicated that the X-F yeasts were present in all <strong>of</strong> the passalid<br />

beetles studied. The most notable geographical difference was that the X-F<br />

members <strong>of</strong> the Scheffersomyces clade, especially S. shehatae and S. stipitis, were<br />

the most common ascomycete yeasts in the gut <strong>of</strong> passalids collected in<br />

Guatemala. By contrast, Scheffersomyces species, except for Scheffersomyces lignicola,<br />

were rare in the gut <strong>of</strong> passalids from Thailand. The most common Thai<br />

yeasts were X-F species in the Spathaspora clade. Based on molecular and other<br />

taxonomic characters, several X-F yeasts were determined to be new species <strong>of</strong><br />

Scheffersomyces and Spathaspora. These results support the conclusion that the<br />

X-F yeasts are closely associated with passalid beetles across continents, although<br />

further studies with more extensive sampling from additional distant regions will<br />

be necessary to understand the nature <strong>of</strong> the yeast associations with passalid beetles.<br />

Urbina, Hector 1 , Sung-Oui Suh 2 , and Meredith Blackwell 1 . 1 Department <strong>of</strong> Biological<br />

2<br />

Sciences, Louisiana State University, Baton Rouge, LA 70803, USA,<br />

Mycology and Botany Program, <strong>America</strong>n Type Culture Collection (ATCC),<br />

Manassas, Virginia 20110, USA. The gut <strong>of</strong> arthropods: a source for isolation<br />

<strong>of</strong> new Trichomonascus species<br />

The gut <strong>of</strong> insects has been a source for the isolation <strong>of</strong> several unrelated<br />

groups <strong>of</strong> yeasts, including members <strong>of</strong> the Candida tanzawaensis and Candida<br />

kruisii clades and species related to Candida albicans (Lodderomyces clade). Recently,<br />

these clades have been significantly enlarged as a result <strong>of</strong> intensive exploration<br />

<strong>of</strong> gut yeast community. The genus Trichomonascus (Trichomonascaceae:<br />

Saccharomycotina) is currently composed <strong>of</strong> 25 yeast species<br />

found primarily on decayed and fresh fungal sporophores or in association with<br />

insects. During our survey <strong>of</strong> gut yeasts, we isolated 93 Trichomonascus strains<br />

from the gut <strong>of</strong> a number <strong>of</strong> beetle families, including Anthribidae, Erotylidae,<br />

Endomychidae, Mycetophagidae, Nitidulidae, Scarabaeidae, Staphylinidae, and<br />

Tenebrionidae collected from Panama and the southeastern USA. Our molecular<br />

and morphological characterization identified the insect-associated Trichomonascus<br />

taxa as Trichomonascus petasosporus and as many as 15 undescribed<br />

species. The new Trichomonascus species are related to Trichomonascus<br />

indianensis and Trichomonascus farinosus previously isolated from fungal<br />

sporophores, as well as Trichomonascus capitulata and Trichomonascus nivea<br />

isolated from plant tissues. The gut <strong>of</strong> fungus-feeding beetles has proven to be a<br />

rich source <strong>of</strong> a number <strong>of</strong> yeasts, including undescribed Trichomonascus species.<br />

U’Ren, Jana M and A Elizabeth Arnold. School <strong>of</strong> Plant Sciences, University <strong>of</strong><br />

Arizona, Tucson, AZ, 85721. Multilocus phylogenetic analysis <strong>of</strong> the Xylariaceae:<br />

what are the roles <strong>of</strong> previously unknown endophytic and endolichenic<br />

fungi?<br />

The Xylariaceae (Sordariomycetes) comprise one <strong>of</strong> the largest and most<br />

diverse families <strong>of</strong> filamentous Ascomycota, with at least 73 accepted genera and<br />

ca. 800 species. Currently recognized species include many facultative saprotrophs<br />

<strong>of</strong> wood, litter, soil and dung, and a few plant pathogens that cause canker<br />

diseases, root rots, and needle blight in agricultural and natural systems. Despite<br />

a strong tradition <strong>of</strong> taxonomic work on Xylariaceae, recent studies illustrate that<br />

an enormous amount <strong>of</strong> undescribed diversity within the family occurs as previously<br />

unstudied endophytic and endolichenic fungi. Previous studies have examined<br />

the evolutionary relationships between described Xylariaceae species and<br />

unknown fungi for a limited number <strong>of</strong> genera using SSU rDNA and ITS rDNA,<br />

but ribosomal genes either lack resolution (i.e., SSU) or are unalignable across diverse<br />

taxa (i.e, ITS), precluding robust phylogenetic analyses at the family level.<br />

We isolated fungi from the interior <strong>of</strong> living leaves, senescent leaves, decaying<br />

leaves, and lichen thalli in five biotically and geographically diverse locations<br />

across North <strong>America</strong>, and sequenced ITS-partial LSU rDNA for 2062 xylariaceous<br />

isolates. Cluster analyses suggested the presence <strong>of</strong> 94 putative species<br />

(based on 95% sequence similarity) and 249 unique genotypes. For one representative<br />

<strong>of</strong> each putative species, we sequenced two protein-coding genes (RPB2<br />

and beta-tubulin), which were aligned with sequence data from 300 described<br />

species <strong>of</strong> Xylariaceae and analyzed using maximum likelihood and Bayesian inference<br />

methods. Preliminary analyses indicate that endophytic and endolichenic<br />

fungi frequently form novel well-supported clades relative to described species.<br />

Some endophytic/endolichenic fungi were found in clades not previously known<br />

to contain symbiotrophic taxa. On-going work will pair phylogenetic analyses<br />

with ancestral state reconstructions to examine the evolution <strong>of</strong> major trophic<br />

modes and the role <strong>of</strong> symbiosis in diversification <strong>of</strong> this important group <strong>of</strong> fungi.<br />

50 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

U’Ren, Jana M and A Elizabeth Arnold. University <strong>of</strong> Arizona, School <strong>of</strong> Plant<br />

Sciences, Tucson, AZ 85719. Surveys at five sites across North <strong>America</strong> reveal<br />

taxonomic and functional differences among fungi from living, senescent,<br />

and fallen leaves<br />

Some fungi recovered from the interior <strong>of</strong> living, asymptomatic leaves<br />

(endophytes) have been isolated from leaf litter, suggesting that many <strong>of</strong> these<br />

symbiotrophs are decomposer fungi with only a transiently endophytic life stage.<br />

However, communities rarely are sampled to statistical completion and surveys<br />

rarely encompass multiple host species and sites. We examined the specificity,<br />

composition, and geographic turnover <strong>of</strong> fungi inhabiting the interior <strong>of</strong> leaves <strong>of</strong><br />

the same host individuals and in the same seasonal cohorts but that represented<br />

three different stages (living, senescent/dead attached to branches, fallen/decomposing).<br />

Analysis <strong>of</strong> 2064 isolates from 3-4 abundant species <strong>of</strong> woody plants in<br />

five temperate and boreal sites consistently revealed that fungal communities<br />

change in response to leaf senescence and incorporation into leaf litter. Although<br />

boundaries among leaf life-stages were blurred by some fungi that occurred in<br />

both living and dead leaves, genotypes that were highly abundant in leaf litter<br />

were seldom found as endophytes. Communities in each leaf stage differed<br />

among hosts and sites, but fungi in dead leaves were less host- and geographically<br />

specific than those in living leaves. Fungi from each leaf class differed in their<br />

relationships with environmental factors, with fungi in dead leaves demonstrating<br />

less strict association with precipitation and growing season duration than endophytes.<br />

On-going work pairing these data with functional traits will inform the<br />

breadth <strong>of</strong> ecological roles <strong>of</strong> endophytes and provide the basis for hypothesis<br />

testing with culture-free methods.<br />

Uribe Valdez, Gilberto and Greg W Douhan. Department <strong>of</strong> Plant Pathology<br />

and Microbiology, University <strong>of</strong> California, Riverside, CA 92521, USA. Hostparasite<br />

relationships between California boletes and bolete-infecting Hypomyces<br />

species.<br />

Fungi in the genus Hypomyces are mycoparasites <strong>of</strong> various fungal<br />

groups, including boletes. Until recently, most studies within this genus primarily<br />

dealt with taxonomic descriptions without much focus on their biology in natural<br />

ecosystems. Currently, there are eight formally described bolete-infecting<br />

Sepedonium anamorphs linked to the Hypomyces teleomorph. Our current objective<br />

is to determine the level <strong>of</strong> host specificity between these parasites and their<br />

hosts using a phylogenetic approach, since previous reports regarding specificity<br />

have been based mostly on morphological observations and a small molecular<br />

dataset. Our long-term goal is to then determine whether we can infer any patterns<br />

<strong>of</strong> co-evolution within this system. As determined in a previous study in California,<br />

two cryptic species were found within H. microspermus and H. chrysospermus<br />

sensu lato, though based on a very limited sample size. We have set out to<br />

corroborate the presence <strong>of</strong> these multiple, phylogenetic species with a multi-gene<br />

phylogeny approach. Currently, sequencing <strong>of</strong> the ITS regions, actin, and elongation<br />

factor genes support the multispecies complex hypothesis, though more sequencing<br />

will be done to further strengthen our findings. Also, two additional<br />

species previously reported to occur in California, H. ampullosporum and H. laevigatum,<br />

will be further analyzed and included in our new host-parasite analyses.<br />

Currently, we have collected over 400 parasitized boletes within California, cultured<br />

the parasites, and isolated DNA from both the hosts and pathogens. Putative<br />

basidiomycete-specific primers are being used to identify and establish host relationships<br />

based on sequencing <strong>of</strong> the variable ITS regions and a partial region <strong>of</strong><br />

the more conserved LSU-rDNA. These host sequences are being used to create<br />

species level phylogenies that will be compared to phylogenies <strong>of</strong> the parasite. We<br />

are finding that many bolete clades are being parasitized by closely related Hypomyces<br />

isolates, and results for all samples in our collection will be presented.<br />

Valent, Barbara. Department <strong>of</strong> Plant Pathology, Kansas State University, Manhattan<br />

KS 66506-5502 USA. Stealth strategies <strong>of</strong> a cereal killer<br />

Magnaporthe oryzae includes host-adapted populations that collectively<br />

threaten rice, wheat and other cereal crops worldwide. Past studies have focused<br />

on understanding development and function <strong>of</strong> the pathogen’s appressorium, the<br />

dome-shaped cell that builds up and focuses enormous pressures to force a tiny<br />

penetration peg through the tough outer plant surface. Now, new understanding<br />

<strong>of</strong> host tissue colonization after penetration has emerged from live cell imaging <strong>of</strong><br />

the fungus expressing fluorescently-labelled components during invasion <strong>of</strong> optically-clear<br />

rice sheath cells. The blast fungus colonizes its host using specialized<br />

intracellular hyphae that successively invade living plant cells. Filamentous hyphae<br />

that enter each rice cell invaginate the host plasma membrane, and then differentiate<br />

into enlarged, bulbous invasive hyphae that are cloaked in a tight coat<br />

<strong>of</strong> plant membrane. Fungal cells directly involved in the morphological switch are<br />

associated with a highly-localized biotrophic interfacial complex (BIC) that appears<br />

to function as the staging center for translocation <strong>of</strong> effectors, pathogen proteins<br />

that hijack host processes, into the cytoplasm <strong>of</strong> the invaded host cell. In-<br />

Continued on following page


deed, some <strong>of</strong> the translocated effectors move ahead into neighboring rice cells<br />

before the fungus enters them, possibly preparing rice cells before invasion. The<br />

thickened invasive hypha again undergoes extreme constriction, forming a tiny<br />

penetration peg-like structure for crossing the plant cell wall into the next cell. The<br />

fungus searches for locations to cross the wall, suggesting that it recognizes and<br />

manipulates the plant’s plasmodesmata, the tiny communication channels between<br />

plant cells, for its cell-to-cell movement. Knowledge important for controlling<br />

a major disease threat to global food security comes back down to understanding<br />

fundamental mechanisms for hyphal polymorphism, for fungal protein<br />

secretion and targeting, and for sensing and adapting to harsh environments.<br />

Vujanovic, Vladimir. University <strong>of</strong> Saskatchewan, Saskatoon, SK, S7N 5A8<br />

Canada. Monitoring Fusarium complex mycelia replacement by mycopathogenic<br />

Sphaerodes using ATPtest, qPCR and HPLC<br />

The biotrophic mycoparasite-host interface and replacement remain poorly<br />

described, yet their understanding is important for effective biocontrol <strong>of</strong><br />

Fusarium plant pathogens. Alcohol percentage test (APT) and real-time PCR<br />

quantification <strong>of</strong> DNA were employed to evaluate the efficiency <strong>of</strong> mycopathogenic<br />

Sphaerodes mycoparasitica mycelia to replace an array <strong>of</strong> pathogenic/toxigenic<br />

Fusaria in co-cultures. Shifts in Fusarium mycelial growth and hydrophobicity<br />

measured (APT) at the interaction zone with S. mycoparasitica indicated<br />

the degree <strong>of</strong> Fusarium mycelia replacement by the mycoparasite (P


associated with Panicum. In 2011, we sampled cultivated P. virgatum in New Jersey<br />

(NJ) and naturally occurring Panicum from New Jersey and Hawaii. Root and<br />

shoot tissue was surface sterilized and plated on acidified malt extract agar. Fungi<br />

were recovered from 79 <strong>of</strong> 1600 root and shoot segments and were identified<br />

based on morphology and DNA sequences <strong>of</strong> the internal transcribed spacer (ITS)<br />

region <strong>of</strong> ribosomal RNA genes. Isolation <strong>of</strong> fungi was twice as high in NJ natural<br />

locations than cultivated. The dominant fungal species isolated from the natural<br />

NJ locations were Fusarium, Bionectria, and Stagnospora. For cultivated NJ<br />

samples, Fusarium, Bionectria, and Cladosporium were the most frequently isolated<br />

fungi. The most prevalent species in all locations was Fusarium oxysporum.<br />

Impact <strong>of</strong> the F. oxysporum strain on the switchgrass host is under investigation.<br />

Wang, Yan 1 , Eric D Tretter 1 , Eric M Johnson 1 , Prasanna Kandel 1 , Robert W<br />

Lichtwardt 2 , and Merlin M White 1 . 1 Department <strong>of</strong> Biological Sciences, Boise<br />

State University, 1910 University Drive, Boise, ID 83725, 2 Department <strong>of</strong> Ecology<br />

& Evolutionary Biology, University <strong>of</strong> Kansas, 1200 Sunnyside Avenue,<br />

Lawrence, KS 66045. Ribosomal RNA gene-based and multi-gene phylogenies<br />

<strong>of</strong> Smittium (Harpellales) and allies-toward unraveling relationships<br />

among early-diverging fungi<br />

Smittium is one <strong>of</strong> the oldest members <strong>of</strong> the Harpellales, a group commonly<br />

referred to as the gut fungi. Gut fungi are endosymbiotic microorganisms<br />

that live in the digestive tracts <strong>of</strong> various Arthropods, worldwide. During the 75<br />

years since the first species, Smittium arvernense, was described, Smittium has<br />

grown to include 81 species. This genus has also helped to advance our understanding<br />

<strong>of</strong> the gut fungi, by serving as a “model” for laboratory studies <strong>of</strong> the fungal<br />

trichomycetes. Many isolates <strong>of</strong> Smittium have been used for ultrastructural,<br />

physiological, host feeding, serological, as well as isozyme, and now ongoing molecular<br />

systematic studies. Previous and current molecular studies have shown<br />

that Smittium is polyphyletic but with consistent separation <strong>of</strong> Smittium culisetae,<br />

one <strong>of</strong> the most common and widespread species, from the remainder <strong>of</strong> Smittium<br />

species. Morphological (zygospore and trichospore shape), molecular (18S and<br />

28S rDNA), immunological, and isozyme evidence were used to establish a new<br />

genus Zancudomyces, and to accommodate Smittium culisetae. A multi-gene<br />

dataset, consisting <strong>of</strong> 18S and 28S rDNA, RPB1, RPB2, and MCM7 translated<br />

protein sequences for Smittium and related Harpellales (Austrosmittium,<br />

Coleopteromyces, Furculomyces, Pseudoharpella, Stachylina and Trichozygospora),<br />

was used for phylogenetic analyses and provided strong support at multiple<br />

levels in the trees generated. The clades and branches <strong>of</strong> the consensus tree<br />

were assessed relative to morphological traits for the taxa <strong>of</strong> interest, including<br />

thallus branching type, holdfast shape, trichospore or zygospore characters as an<br />

aid to inform the taxonomy and eventual systematic revisions and reclassification.<br />

Wang, Zheng 1 , Nina Lehr 1 , Francesc Lopez-Giraldez 1 , Marta Farré 2 , Frances<br />

Trail 3 , and Jeffrey P Townsend 1 . 1 Department <strong>of</strong> Ecology and Evolutionary Biology,<br />

Yale University, New Haven, CT06520, USA, 2 2Departament de Biologia<br />

Cellular, Fisiologia i Immunologia. Universitat Autònoma deBarcelona. Campus<br />

UAB, 08193, Cerdanyola del Vallès, Barcelona, Spain, 3 Department <strong>of</strong> Plant<br />

Biology, Department <strong>of</strong> Plant Pathology, Michigan State University, East Lansing,<br />

MI 48824, USA. Multidimensional regulation in transcriptome during<br />

sexual development in Neurospora crassa revealed with RNA sequencing<br />

Fungi exhibit a huge diversity in morphology and ecology. The genetic<br />

basis <strong>of</strong> morphological development in sexual reproduction <strong>of</strong> multi-cellular<br />

fungi, including the model organism Neurospora crassa, has not been fully investigated<br />

with genome-wide approaches. RNA were sampled for eight time<br />

points cross perithecial development, including two points right before and after<br />

crossing, for which mat A protoperithecia were fertilized with mat a conidia.<br />

Transcription pr<strong>of</strong>iles for 9717 genes were obtained for all eight time-points using<br />

multi-targeting priming and for five time-points using random hexamer priming.<br />

The majority <strong>of</strong> the genome showed continually up- or down- regulated expression<br />

patterns during the later perithecial development stages after 72 and 96 h.<br />

Functionally unclassified proteins were counted for most up-regulated genes,<br />

while genes were enriched with different functions for various down-regulated<br />

patterns. We observed a significant increase in transcription for mat a-1 and<br />

pheromone precursor ccg-4 during perithecial development. In addition, expression<br />

<strong>of</strong> genetic markers for morphological traits in perithecium, ascus, and ascospore,<br />

and for development traits in meiosis <strong>of</strong>ten showed multiple peaks during<br />

the experiment. Expression patterns for genes involved in Meiotic Silencing<br />

and Repeat Induced Point Mutation, and for a gene similar to stc-1, were correlated<br />

and correspondent to morphological development. Interesting transcription<br />

pr<strong>of</strong>iles were also observed for genes involved in heterokaryotic and vegetative<br />

incompatibility, in the heterotrimeric G protein signaling systems, and in the MAP<br />

kinase signaling pathways. Knockouts <strong>of</strong> these regulatory genes are being studied<br />

for their phenotypes in sexual development <strong>of</strong> N. crassa.<br />

Weete, John D 1 , Maritza Abril 2 , and Meredith Blackwell 2 . 1 Auburn Research<br />

& Technology Foundation, Auburn University, Auburn, Alabama 36832 USA,<br />

52 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

2<br />

Department <strong>of</strong> Biological Sciences, Louisiana State University, Baton Rouge,<br />

Louisiana 70803 USA. Biochemical characters support fungal lineages<br />

In the mid 20 th Century, prominent mycologists began to express the<br />

view that fungal biologists had accepted the “strong and growing opinion ... that<br />

the fungi may have originated from animal-like forms.” Before there were DNA<br />

sequences, morphological and biochemical characters (e.g., cell wall components,<br />

flagellation, site <strong>of</strong> meiosis, carbohydrate storage, lysine synthesis, membrane<br />

sterols) were important to separate fungi from other organisms and define monophyletic<br />

groups. The characters have been useful in establishing the degree <strong>of</strong> relatedness<br />

among organisms, but they have supported lineages that later were validated<br />

by DNA analyses. On closer inspection some <strong>of</strong> the characters are more<br />

variable than were appreciated previously. For example the type <strong>of</strong> sterol, previously<br />

thought to be ergosterol for all fungi, actually varies qualitatively within the<br />

fungi. Although ergosterol is the major sterol in the vast majority <strong>of</strong> Dikarya and<br />

certain zygosporic fungi (Mucorales, Zoopagales, Dimargaritales), ergosterol is<br />

not known in all fungi. In fact, end products <strong>of</strong> other sterol pathways (cholesterol,<br />

24-methyl cholesterol, 24-ethyl cholesterol, and brassicasterol) occur as major<br />

sterols among the lineages. For example members <strong>of</strong> Pucciniomycota, the earliest<br />

diverging Basidiomycota, contain 24-ethyl cholesterol; the few zoosporic fungi<br />

tested (Chytridiomycota, Blastocladiomycota) contain a variety <strong>of</strong> sterols with<br />

double bonds only in the C-5 position such as cholesterol and 24-ethyl cholesterol,<br />

and Zygosporic groups outside <strong>of</strong> Mucorales, Zoopagales, and Dimargaritales,<br />

also lack ergosterol and contain a variety <strong>of</strong> C28 sterols. Several studies have used<br />

ergosterol as a measure <strong>of</strong> fungal biomass in the soil, but the method omits arbuscular<br />

mycorrhizal fungi (Glomeromycota), which contain 24-ethyl cholesterol<br />

as the major membrane sterol. Taxon sampling is poor in certain groups, but biochemical<br />

characters hold great promise for looking at lineages as genomics information<br />

becomes readily available.<br />

Williams, Gwendolyn C 1 , Matthew E Smith 2 , Terry W Henkel 3 , and Rytas J<br />

Vilgalys 1 . 1 Department <strong>of</strong> Biology, Duke University, Durham, NC 27708, 2 Department<br />

<strong>of</strong> Plant Pathology, University <strong>of</strong> Florida, Gainesville, FL 32611, 3 Department<br />

<strong>of</strong> Biological Sciences, Humboldt State University, Arcata, CA 95521.<br />

Ectomycorrhizal fungi associated with seedlings <strong>of</strong> a monodominant tree in<br />

the Neotropics<br />

The ectomycorrhizal (EM) leguminous canopy tree Dicymbe corymbosa<br />

forms monodominant stands in the tropical forests <strong>of</strong> the central Guiana Shield region.<br />

Within these stands, D. corymbosa maintains a large bank <strong>of</strong> shade-tolerant<br />

seedlings that result from mast fruiting events. It has been proposed that common<br />

mycorrhizal networks support Dicymbe seedling recruitment and facilitate persistent<br />

monodominance, and that an absence <strong>of</strong> appropriate symbionts may inhibit<br />

these seedlings from successfully establishing in nearby mixed forests <strong>of</strong> AM-associated<br />

trees. Little is known, however, about the actual fungi forming mycorrhizae<br />

with these seedlings. Using molecular methods we examined EM fungi associated<br />

with roots <strong>of</strong> D. corymbosa seedlings within monodominant stands and<br />

along transition zones into AM-dominated mixed forest. Seedlings within monodominant<br />

stands hosted significantly higher EM fungal species diversity than<br />

those in transition zones, but no effect <strong>of</strong> distance from parent tree was detected<br />

in either location. EM fungi commonly recovered on seedlings included members<br />

<strong>of</strong> the Clavulinaceae, Inocybaceae, Russulaceae, and Thelephoraceae, all <strong>of</strong><br />

which have been consistently found on adult D. corymbosa roots. The Boletaceae,<br />

however, were notably under-represented on seedlings. These preliminary results<br />

will be discussed in relation to high-throughput environmental sequencing <strong>of</strong> soils<br />

from the monodominant stands and transition zones.<br />

Williams, Tabitha F 1 , Robert McCleery 2 , Rod McClanahan 3 , and Andrea Porras-Alfaro<br />

1 . 1 Biology Department, Western Illinois University, Macomb, IL<br />

61455, 2 Wildlife Ecology and Conservation PO Box 110430, University <strong>of</strong> Florida,<br />

Gainesville, FL 32611, 3 United States Forest Service, Shawnee National Forest,<br />

Harrisburg, IL 62946. Comparison <strong>of</strong> fungal communities associated with<br />

different bat species in Southern Illinois<br />

Bats are indispensable to ecosystem stability; they play an important role<br />

in the control <strong>of</strong> insect populations and plant pollination. White Nose Syndrome<br />

(WNS), caused by the fungus Geomyces destructans, is a new disease infecting<br />

bats in the USA and Canada. WNS was first reported in 2006 in New York and<br />

has since then spread across the eastern and central United States where it had<br />

made its way to the Illinois borders. The major objective <strong>of</strong> this research was to<br />

identify, characterize and compare psychrophilic (“cold loving”) fungal communities<br />

associated with different bat species in southern Illinois. Bats were trapped<br />

using a harp trapping technique following USGS National Wildlife Center protocols.<br />

Swabs were taken from the bat wings and were inoculated in situ on petri<br />

plates in MEA (Malt Extract Agar) with antibiotics. Samples were incubated at<br />

6°C to select for psychrophilic fungi. We evaluated seven different bat species;<br />

approximately 20 to 30 colonies were obtained in each petri dish, with an average<br />

Continued on following page


<strong>of</strong> four unique morphospecies per swabbed- area (2-3cm 2 ). From the swab plates<br />

about 300 psychrotolerant pure cultures were obtained. DNA extractions, PCR<br />

and sequencing reactions were used to identify fungal cultures. Fungal communities<br />

were dominated by Ascomycota followed by Zygomycota and Basidiomycota.<br />

Common fungi in bat wings include Penicillium corylophilum, Cladosporium<br />

spp, Epicoccum, and Alternaria. Most abundant species were isolated multiple<br />

times from different bat species and caves. This research will increase our knowledge<br />

about all fungal communities associated with bats that are endangered and<br />

currently facing local extinctions.<br />

Wilson, Andrew W 1 , Norman Wickett 1 , Paul Grabowski 2 , Jeremie Fant 1 ,<br />

Justin Borevitz 2,3 , and Gregory M Mueller 1 . 1 Chicago Botanic Garden, Plant<br />

Conservation Science, 1000 Lake Cook Rd, Glencoe, IL 60022, 2 University <strong>of</strong><br />

Chicago, Ecology and Evolution, 1101 E 57th Street, Chicago, IL 60<strong>63</strong>7, 3 Australian<br />

National University, Research School <strong>of</strong> Biology, Canberra, Australia.<br />

Using new techniques in genotyping by sequencing to rejuvenate population<br />

genetic studies in fungi.<br />

Challenges in obtaining sufficient numbers <strong>of</strong> informative population<br />

level markers has been an impediment to understanding the population genetics<br />

<strong>of</strong> fungi. Recent developments in high-throughput technologies have led to the innovation<br />

<strong>of</strong> new methods in genotype by sequencing (GBS) analysis. Restriction<br />

site associated DNA (RAD) sequencing is a relatively new technology that allows<br />

for identification <strong>of</strong> single-nucleotide polymorphisms (SNPs) by sequencing<br />

many small fragments in highly variable regions <strong>of</strong> the genome resulting from a<br />

restriction digest. Restriction sites are fairly conserved within a species, and RAD<br />

has been demonstrated as an effective way to generate loci from which SNPs can<br />

be identified for population genetic analysis. We tested the effectiveness <strong>of</strong> a<br />

novel GBS approach on 18 samples <strong>of</strong> the ectomycorrhizal species Laccaria bicolor.<br />

Initial data collection (one multiplexed lane <strong>of</strong> paired-end HiSeq 2000) produced<br />

over 64 million 101 bp reads. Preliminary analysis <strong>of</strong> all samples identified<br />

13,528 SNPs consisting <strong>of</strong> 1919 multi-allelic loci. A comparison <strong>of</strong> Illinois (3<br />

samples) and Wisconsin (2 samples) populations found 268 heterozygous loci,<br />

which produced an FST <strong>of</strong> 0.3229. Despite the small sample size, these preliminary<br />

results demonstrate the ability <strong>of</strong> the GBS approach to produce the data necessary<br />

for studying Agaricomycete populations. The results <strong>of</strong> a comprehensive<br />

analysis <strong>of</strong> the data generated in this pilot study as well as further discussion on<br />

the potential <strong>of</strong> these methods for fungal population genetic research will be presented.<br />

Wilson, Nathan J 1 , Kathryn M Dunn 2 , Han Wang 2 , and Deborah L McGuinness<br />

2 . 1 Marine Biological Laboratory, Center for Library and Informatics, 7<br />

MBL St., Woods Hole, MA 02543, 2 Rensselaer Polytechnic Institute, 110 Eighth<br />

Street, Troy, NY USA 12180. Application <strong>of</strong> semantic technology to define<br />

names for fungi<br />

The need for well-defined, persistent descriptions <strong>of</strong> taxa that can be accurately<br />

interpreted by computers is becoming increasingly clear. The goal <strong>of</strong> this<br />

work is to develop named descriptions <strong>of</strong> Fungi that enable automated reasoning<br />

by computers. We encode these descriptions using the Web Ontology Language<br />

(OWL). The initial target audience is field mycologists using the Mushroom Observer<br />

website, who range from pr<strong>of</strong>essional scientists to beginning mushroom<br />

enthusiasts. We describe our mycology ontology along with a transparent, community-based<br />

ontology evolution process. The ontology was designed to focus on<br />

properties that can be observed in the field, but the framework is proving to be<br />

suitable for microscopic, chemical, and genomic properties as well. Concrete examples<br />

are provided where field mycologists need names for groups <strong>of</strong> similarlooking<br />

Fungi that are known to belong to different species, and where our approach<br />

can significantly increase the precision <strong>of</strong> information recorded by the<br />

observer. Such a system is important for enabling the field mycologist to make<br />

more meaningful contributions to the modern scientific literature. In addition, the<br />

resulting ontology and descriptions provides a foundation for consistent, unambiguous,<br />

computational representations <strong>of</strong> Fungi. Finally, we expect that such a<br />

system will enable more people to become active field mycologists by providing<br />

a more robust way to document field observations and connect those observations<br />

with information about similar fungi.<br />

Wong, Valerie l, Christopher Schwebach, and Georgiana May. Department <strong>of</strong><br />

Ecology, Evolution, and Behavior, 100 Ecology Building, 1987 Upper Buford<br />

Circle, University <strong>of</strong> Minnesota, Saint Paul, MN 55108. Effects <strong>of</strong> heating on the<br />

frequency and diversity <strong>of</strong> endophytic fungi in boreal Picea and Populus<br />

Fungal endophytes, which live in plant tissue without causing obvious<br />

signs <strong>of</strong> disease, constitute an understudied, hyperdiverse group <strong>of</strong> organisms that<br />

have been implicated in mediating interactions between plants and herbivores,<br />

other fungi, and abiotic factors. This study examines the effect <strong>of</strong> heating on endophytes.<br />

We sampled leaves from Picea glauca and Populus tremuloides in the<br />

B4Warmed experiment, located at the Cloquet Forestry Center, which examines<br />

how Minnesota trees respond to climate warming. Plots <strong>of</strong> seedlings are at ambi-<br />

ent temperature or heated to 1.8°C or 3.8°C above ambient, and these heat treatments<br />

are replicated in closed and open canopy. Over 1,500 leaf samples were<br />

collected from each tree species and plated for culturing. We find a relatively low<br />

isolation frequency in both P. glauca (2%) and P. tremuloides (1.8%). We also<br />

found a comparable frequency <strong>of</strong> non-filamentous isolates (2.4% in P. glauca and<br />

1.3% in P. tremuloides). Slightly more endophytes were found in heat treatment<br />

plots. To our knowledge, our efforts represent the first characterization <strong>of</strong> the endophyte<br />

community in Minnesota’s boreal forest and the repercussions <strong>of</strong> climate<br />

change on that community.<br />

Wu, Tiehang 1 and Dan Chellemi 2 . 1 Department <strong>of</strong> Biology, Georgia Southern<br />

University, Statesboro, GA 30460, 2 USDA-ARS-USHRL, Fort Pierce, FL<br />

34945. Molecular approaches pr<strong>of</strong>iling <strong>of</strong> soil fungal diversity and composition<br />

under different land and crop management systems<br />

Diversity and composition <strong>of</strong> soil and root colonizing fungi were examined<br />

in tomato (Lycopersicum esculentum) plots subjected to diverse land and<br />

crop management systems. Culture-dependent colony counting was used to identify<br />

communities <strong>of</strong> fungi colonizing roots, and length heterogeneity polymerase<br />

chain reaction (LH-PCR) analysis <strong>of</strong> internal transcribed spacer (ITS) pr<strong>of</strong>iles was<br />

used to characterize soil fungal communities. Diversity <strong>of</strong> soil fungi in bahiagrass<br />

(Paspalum notatum var. notatum ‘Argentine’) or undisturbed weed fallow plots<br />

was significantly lower when compared to diversity in organically managed plots.<br />

However, the diversity <strong>of</strong> root colonizing fungi was significantly lower in organic<br />

plots when compared to the bahiagrass or weed fallow plots. Multivariate<br />

analysis <strong>of</strong> root colonizing fungi and genetic ITS-1 pr<strong>of</strong>iles <strong>of</strong> soil fungi both indicated<br />

a higher degree <strong>of</strong> similarity among fungal communities in weed fallow<br />

and bahiagrass plots. Soil fungal communities in organically managed plots displayed<br />

a high degree <strong>of</strong> similarity to each other and were unique when compared<br />

to communities in other land management systems. A dominant 341-bp amplicon<br />

was identified in all soil fungal communities except those from organic plots. The<br />

amplicon was identified as Fusarium oxysporum by cloning and sequence analysis<br />

and confirmed by the LH-PCR amplicon size and sequences for a known<br />

Fusarium oxysporum isolate. Fusarium oxysporum dominated all communities <strong>of</strong><br />

root colonizing fungi except those subjected to organic management practices.<br />

This study addresses the benefits <strong>of</strong> integrating molecular approaches to pr<strong>of</strong>ile<br />

whole soil fungal community with a correlation to target the functional subset <strong>of</strong><br />

root colonizing fungi.<br />

Wu, Yuan. Chang Bai Road 155, Chang Ping District, Beijing, China. Yeast infections<br />

in an intensive care unit <strong>of</strong> a general hospital in Beijing<br />

Hospital infections caused by pathogenic yeasts has increased in recent<br />

years. In the past year, we collected 100 yeasts strains from different medical departments<br />

in a general hospital in Beijing. Thirty-one yeast strains were isolated<br />

from the intensive care unit (ICU) from blood, urine, feces and drainage. Identification<br />

<strong>of</strong> the 31 strains revealed that 13 were Candida tropicalis, 7 were C. albicans,<br />

4 were C. glabrata and C. parasilosis, 2 were Pichia sp. and 1 was Saccharomyces<br />

cerevisiae. In this study, C. tropicalis was the predominant pathogen<br />

isolated from ICU patients, which further confirms that non-albicans Candida infections<br />

in hospitals are increasing.<br />

Young, Darcy 1 , Rachael Martin 1 , James Rice 2 , Dimitrios Floudas 1 , Igor V<br />

Grigoriev 3 , and David S Hibbett 1 . 1 Biology Department, Clark University,<br />

Worcester, MA, 01610, 2 School <strong>of</strong> Engineering, Brown University, Providence,<br />

RI, 02912, 3 US Department <strong>of</strong> Energy Joint Genome Institute, Walnut Creek,<br />

CA, 94598. Analyzing the role <strong>of</strong> white-rot fungi in bioremediation using<br />

transcriptome and degradation assays following growth on Number 6 fuel oil<br />

The extracellular enzymes that white-rot fungi secrete during lignin decay<br />

have been proposed as promising agents for oxidizing persistent pollutants like<br />

polycyclic aromatic hydrocarbons, synthetic dyes, explosives, and pesticides. We<br />

investigated the abilities <strong>of</strong> the white-rot fungi Punctularia strigoso-zonata, Irpex<br />

lacteus, Trichaptum biforme, Phlebia brevispora, Trametes versicolor, and Pleurotus<br />

ostreatus to degrade Number 6 fuel oil in wood sawdust cultures. Our goals<br />

are to advise the bioremediation efforts underway at a brownfield redevelopment<br />

site on the Blackstone River in Grafton, Massachusetts and to contribute to a<br />

broader understanding <strong>of</strong> the mechanisms <strong>of</strong> decay in white-rot fungi. We have<br />

preliminary results from gas chromatography-mass spectrometry analyses suggesting<br />

that Irpex lacteus degrades a significant portion <strong>of</strong> the compounds in No.<br />

6 oil. Punctularia strigoso-zonata is also the subject <strong>of</strong> transcriptome pr<strong>of</strong>iling to<br />

determine if there are genes that are differentially regulated during growth in the<br />

presence <strong>of</strong> No. 6 oil on two different wood substrates. We plan to characterize<br />

the genes in the four sequenced transcriptomes using the unpublished annotated<br />

genome <strong>of</strong> P. strigoso-zonata. A clearer understanding <strong>of</strong> the decay mechanisms<br />

<strong>of</strong> P. strigoso-zonata could streamline the use <strong>of</strong> other white-rot fungi in bioremediation.<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 53


Young-Cheol, Ju, Lee Yun-Hae, and Jang Myoung-Jun. Mushroom Research<br />

Institute, Gyeonggido Agricultural Research & Extension Services 430-8 Sam-Ri<br />

Silchon-Eup Gwangju-Si Gyeonggi-Do, Postal code : 464-873. Utility <strong>of</strong> Albasia<br />

sawdust for cultivating Pleurotus ostreatus in bottles<br />

In this study, we attempted to find substitute materials for cottonwood<br />

sawdust used in the cultivation <strong>of</strong> Pleurotus ostreatus in bottles. Chemical characteristics<br />

<strong>of</strong> mixed substrates substituting albasia sawdust for cottonwood sawdust<br />

were not different significantly. Tests for 184 pesticides (Acrinathrin, etc.)<br />

showed that none were detected in the sawdust. The incubation period was shorter<br />

using 50% albasia sawdust compared to 100% albasia sawdust. The yield <strong>of</strong><br />

fruit-bodies was similar to the control and those obtained with 50% albasia sawdust.<br />

However, the 50% albasia sawdust treatment was higher in bio-efficiency<br />

than the control. Therefore, it is suggested that the 50% albasia sawdust treatment<br />

be substituted for cottonwood sawdust for cultivation <strong>of</strong> P. ostreatus.<br />

Zelski, Steven E 1 , Andrew N Miller 2 , and Carol A Shearer 1 . 1 Department <strong>of</strong><br />

Plant Biology, University <strong>of</strong> Illinois, 505 South Goodwin Ave., Urbana, IL<br />

61801, 2 Illinois Natural History Survey, University <strong>of</strong> Illinois, 1816 South Oak<br />

St., Champaign, IL 61820. Environmental sampling as a means <strong>of</strong> rapid identification<br />

<strong>of</strong> freshwater fungal assemblages on submerged woody debris<br />

Quickly identifying the members <strong>of</strong> a fungal community colonizing submerged<br />

woody debris is a goal for those interested in studying aquatic fungi. Traditional<br />

methods are both time consuming and require expert knowledge <strong>of</strong> freshwater<br />

fungal taxonomy, thus a molecular identification method is desirable to<br />

address these issues. While environmental sampling has been done on various<br />

other substrates, no previous work has been performed on submerged woody debris.<br />

To fill this void a sampling protocol for submerged woody debris was developed<br />

to follow changes in the fungal community on submerged balsa wood<br />

baits over a seven month time period. Balsa wood baits were attached to bricks<br />

and set out in two transects in an Illinois stream. One sample from each transect<br />

was retrieved each month <strong>of</strong> the study and water temperature and pH were recorded.<br />

Water temperature ranged from 2.2-19.5 C and pH ranged from 8.0-10.0.<br />

Whole community fungal DNA was extracted from the wood samples, the Internal<br />

Transcribed Spacer (ITS) <strong>of</strong> the nuclear rDNA was amplified, clone libraries<br />

were constructed, and individual clones were sequenced. Sequences were then<br />

sorted into operational taxonomic units and BLAST searches performed in Gen-<br />

Bank. In total, 25 taxa were recovered over the course <strong>of</strong> the study. Species diversity<br />

was found to be lower on balsa wood baits than on naturally occurring substrates,<br />

which may reflect differences in substrate quality. This study illustrates<br />

that not only is it possible to rapidly assess the species present on submerged<br />

woody debris, but that there is also the potential to recover a greater diversity <strong>of</strong><br />

species using this method.<br />

Zelski, Steven E 1 , Carol A Shearer 1 , Huzefa A Raja 3 , and Andrew N Miller 2 1<br />

.<br />

Department <strong>of</strong> Plant Biology, University <strong>of</strong> Illinois, 505 S. Goodwin Ave., Urbana,<br />

IL 61801, 2 Illinois Natural History Survey, University <strong>of</strong> Illinois, 1816<br />

South Oak St., Champaign, IL 61820, 3 Department <strong>of</strong> Chemistry and Biochemistry,<br />

University <strong>of</strong> North Carolina Greensboro, 457 Sullivan Science Building,<br />

Greensboro, NC 27402-6170. Species richness and distribution patterns <strong>of</strong><br />

freshwater ascomycetes along an altitudinal gradient in the Peruvian Andes<br />

We sampled freshwater ascomycetes along an altitudinal gradient in the<br />

Peruvian Andes to determine their spatial distribution patterns along a climatic<br />

gradient. Previous, but as yet unpublished work has shown a near complete<br />

turnover <strong>of</strong> species along a latitudinal gradient reaching from the sub-Arctic to the<br />

tropics and one might expect a similar pattern to emerge with elevation, as has<br />

been shown for terrestrial plants and aquatic invertebrates. Submerged woody debris<br />

was collected from a variety <strong>of</strong> aquatic habitats including palm swamps,<br />

ponds, streams and rivers along with environmental data such as water temperature,<br />

pH, dissolved oxygen content and GPS location. For data analyses, sites<br />

were categorized as low, medium, or high elevation along an altitudinal gradient<br />

ranging from 218-3870 m. Approximately 30 pieces <strong>of</strong> decomposing plant debris<br />

were collected at each site and examined in the laboratory for the presence <strong>of</strong> fun-<br />

54 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

gal fruiting structures. Data for this abstract is based on two collecting trips, one<br />

in May and the other in September and October <strong>of</strong> 2010. Over 1250 individual<br />

fungal specimens have been documented, representing nearly 300 species. Preliminary<br />

analysis <strong>of</strong> the distribution data suggests that the three elevational levels<br />

differ in their community assemblages and that species turnover occurred along<br />

the altitudinal gradient. Species richness was highest at mid elevation (153<br />

species), followed by low elevation (95 species) and high elevation (39 species).<br />

Low and mid elevation showed the highest degree <strong>of</strong> similarity (.481 Sørensen<br />

similarity index), while both low and mid elevations showed low similarity to<br />

high elevation (.285 and .2<strong>63</strong> Sørensen similarity index, respectively). Comparison<br />

<strong>of</strong> the Peru species with a summary list <strong>of</strong> species from several inventories<br />

from another tropical region, namely Thailand, reveals a Sørensen similarity<br />

index <strong>of</strong> .152 suggesting there are pantropical species, but differences in longitudinal<br />

assemblages do exist.<br />

Zhang, Ning. Dept. <strong>of</strong> Plant Biology and Pathology, Rutgers University, New<br />

Brunswick, NJ 08901. Barcodes and a MycoCHIP - a case study on turfgrass<br />

pathogen barcoding and diagnosis<br />

The internal transcribed spacer (ITS) region <strong>of</strong> ribosomal RNA genes is<br />

one <strong>of</strong> the most commonly sequenced loci in fungi and has been declared as the<br />

<strong>of</strong>ficial fungal barcode. Usually ITS sequences allow differentiation <strong>of</strong> fungi at<br />

the species level, which makes them ideal for designing species specific probes in<br />

molecular detection methods such as DNA diagnostic array. The diagnostic array<br />

tool <strong>of</strong>fers a fast, culture-independent alternative for the detection <strong>of</strong> microbes<br />

from field samples. It supports simultaneous detection and differential identification<br />

<strong>of</strong> multiple target species in a single assay. Our current objective is to develop<br />

a highly sensitive Turf PathoCHIP for rapid detection <strong>of</strong> known and emerging<br />

turfgrass pathogens, based on ITS sequences <strong>of</strong> fungi. The development <strong>of</strong> this<br />

product will allow for efficient pathogen identification resulting in more effective<br />

disease management, improved plant health and potential reductions in pesticide<br />

usage. As the bacterial PhyloChip, this technology can be extended to form a<br />

“MycoCHIP”, which may cover fungi <strong>of</strong> all known diversity, in order to detect<br />

and identify fungi in other habitats on earth.<br />

Zimmerman, Kolea C 1 , Daniel Levitis 2 , and Anne Pringle 1 . 1 Department <strong>of</strong><br />

Organismic<br />

2<br />

and Evolutionary Biology, Harvard University, Cambridge MA,<br />

Max Planck Institute for Demographic Research, Rostock, Germany. High<br />

throughput analyses <strong>of</strong> early life history traits <strong>of</strong> Neurospora crassa<br />

Many organisms experience high mortality during initial growth. This is<br />

especially relevant for most fungi because <strong>of</strong> the large ratio between propagules<br />

produced and the number <strong>of</strong> those propagules that germinate and grow into a mature<br />

mycelium. In this study, we developed flow cytometry methods that can be<br />

used to determine the fine scale germination dynamics <strong>of</strong> Neurospora crassa<br />

conidia and ascospores. Specifically, we determined germination dynamics <strong>of</strong><br />

conidia harvested from eight genetically distinct WT strains and we also optimized<br />

conditions for the cytometry <strong>of</strong> ascospores. Conidia and ascospores were<br />

germinated in liquid culture and sampled at intervals. Aliquots were fixed and nucleic<br />

acids <strong>of</strong> the germinating spores were fluorescently labeled. Processed samples<br />

were measured with a BD Fortessa flow cytometer using a 488 nm laser for<br />

all parameters. Spearman rank correlations showed that side-scatter-width and<br />

fluorescence-width were the most informative parameters for germination over<br />

time. Increased nucleic acid content preceded physical growth and nucleic acid<br />

content was the most variable growth trait. This indicates that nucleic acid content<br />

and rate <strong>of</strong> change are informative parameters for further quantitative genetic and<br />

heritability analyses. Germination rates and percent germination <strong>of</strong> conidia differed<br />

significantly amongst the wild isolates tested. Finally, we developed a<br />

framework to test the effects <strong>of</strong> parent relatedness on the viability <strong>of</strong> meiotically<br />

produced <strong>of</strong>fspring. We calculated precise genetic distances between environmental<br />

isolates <strong>of</strong> N. crassa using publicly available RNAseq data and devised<br />

crosses to represent the range <strong>of</strong> calculated distances. Our research will facilitate<br />

further experimental work on the evolutionary demography <strong>of</strong> early life history<br />

traits in the Fungi.


MYCOLOGIST’S BOOKSHELF<br />

We have two reviews for this issue, and I’ve added a few new titles to the list <strong>of</strong> available books below.<br />

Just as a reminder, I’ve adopted a new process for getting books to you, whereby books will be “drop-shipped”<br />

directly from publisher to reviewer; this will streamline the process at my end and will save the <strong>Society</strong> the<br />

shipping expense. Also <strong>of</strong> note, and perhaps a trend: one <strong>of</strong> the publishers whose books we frequently review<br />

is adopting a new policy that currently only applies to some <strong>of</strong> their books: the review will have to be done<br />

using an online version (though not a pdf or something for a Kindle or e-reader); once they receive a copy <strong>of</strong><br />

the published review, they’ll ship the reviewer the hard copy. This will be the case with the Springer “Laboratory<br />

Protocols in Fungal Biology” by Gupta et al.<br />

If you would like to review a book or CD, please contact me (robert.marra@ct.gov). A book goes to the<br />

first person requesting it, and I ask that you get your reviews to me in a reasonably timely manner. Also, if you<br />

know <strong>of</strong> a newly published book that might be <strong>of</strong> interest to mycologists, please let me know so I can request<br />

it from the publisher.<br />

—Bob Marra<br />

Books in Need <strong>of</strong> Reviewers<br />

• Biodiversity in Dead Wood. 2012. Juha Siitonen, Bengt Gunnar<br />

Jonsson. Cambridge University Press, Cambridge, UK.<br />

www.cambridge.org. ISBN: 9780521717038. 524 pp, 92 b/w<br />

illus. 21 tables. Price: £38.00 (paperback). New this Issue.<br />

• Fungi <strong>of</strong> Serbia and Western Balkans [Gljive Srbije I Zapadnog<br />

Balkana]. 2009. Branislav Uzelac (photographs by<br />

Goran Milosevic). BGV Logik, Belgrade, Serbia. Branislav<br />

Uzelac. 2009. BGV Logik, Serbia (english@glijvari.org.rs) or<br />

(goran.milosevic@poducavanje.co.rs. ISBN 978-86-912677-0-<br />

4. 464 pp., ca. 1200 color photographs. Price 120 €. New this<br />

Issue.<br />

• Cytology and Plectology <strong>of</strong> the Hymenomycetes. 2012 (2 nd revised<br />

edition). Heinz Clémençon. Schweizerbart Science Publishers.<br />

Stuttgart. www.schweizerbart.de. ISBN 978-3-443-<br />

50037-5. 520 pp, <strong>63</strong>6 figures, 12 tables. Price: €98.00<br />

(paperback). New this Issue.<br />

• The Mycota, Vol. 10: Industrial Applications, 2 nd ed. 2011.<br />

Martin H<strong>of</strong>richter (Ed.). Karl Esser (Series Ed.). Springer-Verlag,<br />

Berlin. www.springer.com. ISBN: 978-3-642-11457-1. 485<br />

pp, 152 illus. Price: $269.00 (hardcover).<br />

• Biomonitoring, Ecology, and Systematics <strong>of</strong> Lichens. Bibliotheca<br />

Lichenologica vol 106. 2011. Scott T Bates, Frank Bungartz,<br />

Robert Lucking, Maria A Herrera-Campos, Angel Zambrano<br />

(Eds.). Borntraeger/Schweizerbart Science Publishers,<br />

Stuttgart. www.schweizerbart.de. ISBN: 978-3-443-58085-8 442<br />

pp, 102 figures, 33 tables, 16 color plates. Price: €109.00 (s<strong>of</strong>tcover).<br />

• Plant Fungal Pathogens: Methods and Protocols. 2012.<br />

Melvin D Bolton, Bart PHJ Thomma (Eds). Part <strong>of</strong> the “Methods<br />

in Molecular Biology” series, v.835. Springer-Verlag, Berlin.<br />

www.springer.com. ISBN: 978-1-61779-5008. 769 pp, 138<br />

illus., 74 in color. Price: $159.00 (hardcover).<br />

• Key For Identification <strong>of</strong> Common Phytophthora Species.<br />

2011. Jean Beagle Ristaino. APS Press, St. Paul, MN. www.apsnet.org.<br />

ISBN: 978-0-89054-397-9. CD-ROM. Price: $269.00<br />

(single user).<br />

• Pollination Biology. 2012. D. P. Abrol. Biodiversity, Conservation,<br />

and Agricultural Production. Springer-Verlag, Berlin.<br />

www.springer.com. ISBN: 978-94-007-1941-5. 792pp, 23 illus.,<br />

18 in color. Price: $209.00 (hardcover).<br />

• Fundamentals <strong>of</strong> Mold Growth in Indoor Environments and<br />

Strategies for Healthy Living. 2011. Olaf CG Adan & Robert<br />

A Samson (Eds). Wageningen Academic Press, Wageningen.<br />

www.wageningenacademic.com. 978-90-8686-135-4. 524 pp.<br />

(hardback). Price: €97.<br />

• Elements <strong>of</strong> Evolutionary Genetics. 2010. Brian Charlesworth<br />

and Deborah Charlesworth. Roberts & Company Publishers,<br />

Greenwood Village, CO. www.roberts-publishers.com/. ISBN:<br />

978-0-9815-1942-5. 768 pp, b&w. Price: $64 (s<strong>of</strong>tcover).<br />

• The Analysis <strong>of</strong> Biological Data. 2009. Michael Whitlock and<br />

Dolph Schluter. Roberts & Company Publishers, Greenwood<br />

Village, CO. www.roberts-publishers.com/. ISBN: 978-0-9815-<br />

1940-1. 704 pp, full color. Price: $80 (hardback).<br />

• Evolution: Making Sense <strong>of</strong> Life. Available August 2012<br />

(copyright year 2013). Carl Zimmer and Douglas J. Emlen.<br />

Roberts & Company Publishers, Greenwood Village, CO.<br />

www.roberts-publishers.com/. ISBN: 978-1-9362-2117-2. 800<br />

pp, printed in four colors. Price: $92 (hardback).<br />

• Laboratory Protocols in Fungal Biology: Current Methods<br />

in Fungal Biology. Avail June 2012. Vijai Kumar Gupta, Maria<br />

Tuohy, Eds. Springer-Verlag, Berlin. www.springer.com. ISBN:<br />

978-1-4614-2355-3. 802 pp, 105 illus., 40 in color. Price:<br />

$279.00 (hardcover). Review Copy online only; hardcover copy<br />

provided following receipt <strong>of</strong> published review.<br />

• Fungal Plant Pathogens (Principles and Protocols Series).<br />

March 2012. CR Lane, P Beales, KJK Hughes (Eds). CABI, Oxfordshire,<br />

UK. www.cabi.org. 978-1-8459-3668-6. 324 pp. Price:<br />

$75.<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 55


Books with Reviewers Assigned<br />

• Waxcap Mushrooms <strong>of</strong> Eastern North <strong>America</strong>. 2012. Alan E.<br />

Bessette, William C. Roody, Walter E. Sturgeon, and Arleen R.<br />

Bessette. Syracuse University Press, Syracuse, NY. www.syracuseuniversitypress.syr.edu.<br />

ISBN: 978-0-8156-3268-9. 192 pp,<br />

167 color, 3 b&w illustrations. Price: $95.00 (hardcover). Reviewed<br />

this issue.<br />

• Biology <strong>of</strong> Lichens – Symbiosis, Ecology, Environmenta Monitoring,<br />

Systematics and Cyber Applications. Bibliotheca<br />

Lichenologica vol 105. 2010. Thomas Nash III, Linda Geiser,<br />

Bruce McCune, Dagmar Triebel, Alexandru M Tomescu, William<br />

Sanders (Eds.). Borntraeger/Schweizerbart Publishers, Stuttgart.<br />

www.schweizerbart.de. ISBN: 978-3-443-58084-1 256 pp, 81 figures,<br />

19 tables. Price: €79.00 (s<strong>of</strong>tcover). Reviewed this issue.<br />

• Practical Guide to Turfgrass Fungicides. 2011. Richard Latin.<br />

APS Press, St. Paul, MN. www.apsnet.org. ISBN: 978-0-89054-<br />

392-4. 280 pp, 115 images, 29 chemical structures. Price: $139.95<br />

(hardcover).<br />

• Coalescent Theory: An Introduction. 2009. John Wakely.<br />

Roberts & Company Publishers, Greenwood Village, CO.<br />

www.roberts-publishers.com/. ISBN: 978-0-9747-0775-4. 352 pp,<br />

b&w. Price: $48 (s<strong>of</strong>tcover).<br />

• The Tangled Bank: An Introduction to Evolution. 2010. Carl<br />

Zimmer. Roberts & Company Publishers, Greenwood Village,<br />

CO. www.roberts-publishers.com/. ISBN: 978-0-9815-1947-0.<br />

394 pp, full color. Price: $48 (hardback).<br />

• Biology <strong>of</strong> Marine Fungi. 2012. Chandralata Raghukumar (Ed).<br />

Part <strong>of</strong> the “Progress in Molecular and Subcellular Biology” series,<br />

v.53. Springer-Verlag, Berlin. www.springer.com. ISBN: 978-3-<br />

642-23341-8. 354 pp, 83 illus., 27 in color. Price: $209.00 (hardcover).<br />

• Tree Thinking: An Introduction to Phylogenetic Biology.<br />

Available July 2012. David Baum and Stacey Smith. Roberts &<br />

Waxcap Mushrooms <strong>of</strong> Eastern North <strong>America</strong><br />

Waxcap Mushrooms<br />

<strong>of</strong> Eastern North <strong>America</strong>.<br />

2012. By Alan E. Bessette,<br />

William C. Roody,<br />

Walter E. Sturgeon, and<br />

Arleen R. Bessette. Syracuse<br />

University Press.<br />

ISBN: 978-0-8156-3268-<br />

9. 179 pages. Cloth<br />

$95.00.<br />

Mushrooms belonging<br />

to the Hygrophoraceae<br />

are some <strong>of</strong><br />

the most visually striking<br />

among the agaric fungi.<br />

Many are brightly colored,<br />

<strong>of</strong>ten easily identified on macroscopic appearance<br />

alone, and consequently some <strong>of</strong> these are routinely included<br />

in general mushroom field guides. Commonly referred to<br />

as “waxcaps,” they are <strong>of</strong>ten encountered in field and forest<br />

56 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Company Publishers, Greenwood Village, CO. www.roberts-publishers.com/.<br />

ISBN: 978-1-9362-2116-5. 400 pp, b&w. Price: $60<br />

(hardback).<br />

• Parasites in Ecological Communities: from Interactions to<br />

Ecosystems. 2011. Melanie J. Hatcher and Alison M. Dunn. Part<br />

<strong>of</strong> the Ecology, Biodiversity and Conservation series. Cambridge<br />

University Press, Cambridge, UK. www.cambridge.org. ISBN:<br />

9780521718226. 464 pp, 113 b/w illus., 7 tables. Price: $60.00 (paperback).<br />

• Forest Health: An Integrated Perspective. 2011. John D. Castello,<br />

Stephen A. Teale (eds.). Cambridge University Press, Cambridge,<br />

UK. www.cambridge.org. ISBN: 9780521766692. 404 pp,<br />

80 b/w illus. 7 tables. Price: £65.00 (hardcover).<br />

• Yeast Research: An Historical Overview. 2011. James A. Barnett,<br />

Linda Barnett. ASM Press, Washington, DC. www.asmpress.org.<br />

ISBN: 978-1-55581-516-5. 392 pp, illus. Price: $159.95<br />

(hardcover).<br />

• Medically Important Fungi: A Guide to Identification, 5 th ed.<br />

2011. Davise H. Larone. ASM Press, Washington, DC. www.asmpress.org.<br />

ISBN: 978-1-55581-660-5. 508 pp, illus, color plates.<br />

Price $109.95 (hardcover).<br />

• The Mycota, Vol. 14: Evolution <strong>of</strong> Fungi and Fungal-Like Organisms.<br />

2011. Stefanie Pöggeler, Johannes Wöstemeyer (Eds.).<br />

Karl Esser (Series Ed.). Springer-Verlag, Berlin.<br />

www.springer.com. ISBN: 978-3-642-19973-8. 345 pp, 60 illus.,<br />

10 in color. Price: $269.00 (hardcover).<br />

• Microbial Bi<strong>of</strong>ilms: Current Research and Applications. 2012.<br />

Gavin Lear, Gillian D. Lewis (Eds.). Caister Academic Press, Norfolk,<br />

UK. www.caister.com. 228 pp. Price: GB £159, US $310<br />

(hardcover).<br />

as saprobes or mycorrhizal symbionts in both tropical and<br />

temperate zones.<br />

This guide is touted as comprehensive coverage to the<br />

waxcaps as they occur in eastern North <strong>America</strong>. That region<br />

is defined as portions <strong>of</strong> Canada and the United States east <strong>of</strong><br />

the Rocky Mountains and is illustrated on a map. The authors<br />

have divided the book into logical portions: a preface, acknowledgements,<br />

and map precede an introductory section<br />

outlining the group covered, attribution to those who have<br />

contributed expertise to waxcap diversity, enumeration <strong>of</strong><br />

macroscopic features, edibility, and possible lookalikes from<br />

other genera. While several genera are almost universally<br />

recognized in the Hygrophoraceae, the authors justify their<br />

acceptance <strong>of</strong> two in this guide, Hygrocybe and Hygrophorus.<br />

In defining the group, the authors give some attention to<br />

the waxy nature <strong>of</strong> the basidiomes as well as microscopic details<br />

(e.g., the orientation <strong>of</strong> the lamellar trama) used to characterize<br />

genera. Descriptions and color images follow and<br />

the guide finishes up with a glossary, bibliography, separate<br />

Continued on following page


indices to common and scientific names, photo credits, and<br />

short bio-sketches <strong>of</strong> the authors.<br />

The main part <strong>of</strong> the book is taken up with 70 pages <strong>of</strong><br />

descriptions followed by slightly more than 160 high quality<br />

color images. The descriptive text is alphabetical by species<br />

(Hygrocybe then Hygrophorus) and for a particular taxon<br />

gives the accepted name, synonymy, if any, page number <strong>of</strong><br />

illustration, common name, concise macroscopic description,<br />

spores size and morphology, followed by observations<br />

on distribution and ecology, a note on edibility, and a commentary<br />

usually giving comparisons to and distinctions from<br />

other species. The color images are mostly confined to an approximately<br />

80 page color plate section where 2, 8.5 × 13 cm<br />

pictures reside on a page; but there are 8 exceptions (pp. 89,<br />

91, 98, 109, 116, 118, 124, 153) where some images are<br />

slightly smaller. Image quality is outstanding, and the majority<br />

<strong>of</strong> the images were captured in natural settings, although<br />

there are few studio images as well. These color plates are<br />

grouped by genus, Hygrocybe then Hygrophorus, and within<br />

those, the arrangement follows a color scheme from white or<br />

pale colored species to those that are more highly pigmented.<br />

There is one subgroup <strong>of</strong> Hygrocybe, called “Chameleons”<br />

for those usually reddish hygrocybes with changeable colors<br />

(e.g., H. conica).<br />

There are no technical identification aids (i.e., dichotomous<br />

or synoptic keys) <strong>of</strong> any kind. If the user is attempting<br />

an identification <strong>of</strong> a waxcap, then the only way to do this is<br />

to compare a collection with an image and/or read a description.<br />

This method may well work for the average to advanced<br />

user, and I suspect that this modus is one employed more<br />

Biology <strong>of</strong> Lichens<br />

Biology <strong>of</strong> Lichens –<br />

Symbiosis, Ecology, EnvironmentalMonitoring,<br />

Systematics and<br />

Cyber Applications.<br />

2010. By Thomas Nash<br />

III, Linda Geiser, Bruce<br />

McCune, Dagmar<br />

Triebel, Alexandru M<br />

Tomescu, William<br />

Sanders (Eds.). Bibliotheca<br />

Lichenologica vol 105.<br />

Borntraeger/Schweizerbart<br />

Publishers, Stuttgart.<br />

256 pp, 81 figures, 19 tables.<br />

ISBN: 978-3-443-<br />

58084-1. Price: €79.00<br />

(s<strong>of</strong>tcover).<br />

Bibliotheca Lichenologica is a series including several<br />

monographs <strong>of</strong> lichens within various taxonomic groups or<br />

from cryptogamically unexplored geographic regions. To<br />

date, the series comprises 107 volumes, the earliest <strong>of</strong> which<br />

date back to the mid-1980s. Biology <strong>of</strong> Lichens – Symbiosis,<br />

Ecology, Environmental Monitoring, Systematics and Cyber<br />

<strong>of</strong>ten than not with general field guides. The authors clearly<br />

do not intend this guide as an authoritative scientific treatment,<br />

but rather as a comprehensive illustrative guide to hygrophori<br />

which they believe fills a gap between general field<br />

guides and monographic treatments. This lack may seem as<br />

a fault to some, but many <strong>of</strong> these taxa are readily identifiable<br />

from macroscopic features.<br />

There are two idiosyncrasies <strong>of</strong> note. In more than one<br />

instance, nomenclatural novelties are provided. For example,<br />

Hygrocybe basidiosa (Peck) on p 17 is <strong>of</strong>fered as a new combination,<br />

but the authors do not provide the original bibliographic<br />

reference even though the basionym is given. Further,<br />

in the observations portion <strong>of</strong> the description, another<br />

comb. nov., Hygrocybe albipes (Peck) is presented without<br />

basionym and without original bibliographic citation. Both<br />

<strong>of</strong> these instances are invalid and that is a pity because proper<br />

nomenclatural compliance would have been easily accomplished.<br />

Secondly, publishers and authors <strong>of</strong> mushroom field<br />

guides always publish a disclaimer about using the book for<br />

determining if an entity is edible. I would then question why<br />

there is a two page section on edibility, and in the descriptive<br />

section, why there is a note on edibility <strong>of</strong> a species. Those<br />

quirks aside, there are quite satisfactory descriptions coupled<br />

with some <strong>of</strong> the best photographs <strong>of</strong> mushrooms that I have<br />

ever seen. The goal <strong>of</strong> providing non-technical, well-illustrated<br />

documentation for waxcaps <strong>of</strong> eastern North <strong>America</strong><br />

is well-done.<br />

—Roy E. Halling<br />

The New York Botanical Garden<br />

rhalling@nybg.org<br />

Applications (Volume 105) includes selected papers related<br />

to the International Association for Lichenology (IAL) 6<br />

symposium held in Asilomar, California in 2008. The first<br />

IAL symposium, held at the University <strong>of</strong> Münster in 1986,<br />

was published in Volume 25 <strong>of</strong> Bibliotheca Lichenologica,<br />

and IAL 3 and IAL 4 were published in Volumes 68 and 82,<br />

respectively. Proceedings for IAL 2 were published in Cryptogamic<br />

Botany, while Folia Cryptogamica Estonia contained<br />

several contributions from IAL 5.<br />

Biology <strong>of</strong> Lichens – Symbiosis, Ecology, Environmental<br />

Monitoring, Systematics and Cyber Applications includes<br />

26 papers compiled into 9 sections. As the title implies, the<br />

papers cover a broad spectrum <strong>of</strong> lichen topics. This volume<br />

includes: holes in the symbiosis literature; current advances<br />

in data networking as related to lichenology; recent studies<br />

using lichens to quantify air pollution; and reports <strong>of</strong> new<br />

species. The remaining third <strong>of</strong> the papers fit roughly into the<br />

Ecology placeholder specified in the book subtitle, yet have<br />

little true topical overlap.<br />

As paper and poster presentations from an international<br />

symposium, each <strong>of</strong> the studies <strong>of</strong>fers novel research, which<br />

contributes to the field <strong>of</strong> lichenology as a whole. The greatest<br />

original strength <strong>of</strong> this compilation, however, is the section<br />

Continued on following page<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 57


on data networking. Significant progress has been made in the<br />

past two decades to query the volumes <strong>of</strong> information amassed<br />

in the literature and in herbaria on lichens. These search tools<br />

(e.g., LIAS, CNALH, RLL and KeyToNature) are freely<br />

available to novices and experts alike, from every corner <strong>of</strong> the<br />

earth where a lichen grows and piques someone’s curiosity.<br />

This section <strong>of</strong> four informative papers provides a tidy description<br />

<strong>of</strong> this work largely lacking from the lichen literature<br />

(as evidenced from RLL and Mattick’s index).<br />

My favorites among this melting pot <strong>of</strong> lichen bioinformation,<br />

however, were the two papers by Tomescu et al.<br />

(2010) and Schwartzman (2010). Tomescu and others’ inventive<br />

methodology to simulate fossilization (i.e., 4 days <strong>of</strong> wet<br />

compression followed by 4 hours <strong>of</strong> heating at 130°C), represents<br />

a novel way to understand the prevalence <strong>of</strong> problematic<br />

cryptogamic fossils or paucity <strong>of</strong> non-vascular plant and<br />

fungal fossils all together. Moreover, their study compliments<br />

Schwartzman’s insights on the environmental triggers for the<br />

lichen symbiosis. I found his paper a refreshing read <strong>of</strong> big<br />

ideas and little data. His hypothesis transported me millions<br />

<strong>of</strong> years into the past, and let me watch a gasping photobiont<br />

clutch its way around its surroundings until finally collapsing<br />

into the hyphae <strong>of</strong> its partner and taking in a deep breath.<br />

58 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

The primary weakness <strong>of</strong> the book is its lack <strong>of</strong> coherence.<br />

Within 256 pages, there were 9 sections, which made<br />

the text read like a journal rather than a book. Only one line<br />

on the inside cover clarifies that this volume was intended to<br />

reflect the variety <strong>of</strong> lichen information presented at IAL 6;<br />

nevertheless, as a single work it is a bit disjointed. Upon<br />

reading the title and subtitle, I had anticipated a general introduction<br />

on lichen biology providing a primer on a variety<br />

<strong>of</strong> topics for new lichenologists. In retrospect, the title aptly<br />

describes the contents <strong>of</strong> this text, yet its generality led me to<br />

anticipate a book suitable for novices. Prolific esoteric lichen<br />

terminology, the journal-like format, and studies narrowly<br />

focusing on one geographic problem or taxonomic group<br />

limit its utility to this audience. Ironically, this text may be<br />

more appropriate for pr<strong>of</strong>essional lichenologists who, in all<br />

likelihood, attended the IAL symposium themselves. So if<br />

you prepare yourself for a journal-like read, you will be rewarded<br />

with gems <strong>of</strong> lichen inquiry and documentation <strong>of</strong><br />

the technological transformation <strong>of</strong> our field.<br />

REMINDER: MSA Directory Update<br />

—Emily A Holt<br />

Department <strong>of</strong> Biology<br />

Utah Valley University<br />

eholt@uvu.edu<br />

MYCOLOGICAL CLASSIFIEDS<br />

Mold and Fungus Testing and Identification Services<br />

Biochallenge tests for ink, micr<strong>of</strong>luidic materials; testing for resistance<br />

<strong>of</strong> materials to fungal invasion. Identification <strong>of</strong> fungal contaminants<br />

in manufactured products. Epifluorescent microbial detection<br />

in deionized water systems, micr<strong>of</strong>luidic devices, medical<br />

fluids, manufactured goods. Identification <strong>of</strong> fungi from buildings,<br />

animal and plant diseases. 10% discount for regular and sustaining<br />

MSA members. Email info@pacificanalytical.com. For more information<br />

see www.pacificanalytical.com<br />

Biological Control, Biotechnology and Regulatory Services<br />

Center for Regulatory Research, LLC specializes in regulatory<br />

permit application services for biological control and biotechnology<br />

organisms/products. Let us evaluate your research discoveries for<br />

commercial potential and environmental impacts. We also <strong>of</strong>fer assistance<br />

with writing proposals for SBIR grant programs (Small<br />

Business Innovation Research) that fund new commercial ventures.<br />

Contact Dr. Sue Cohen by email (sdcohen@regresearch.com) or by<br />

phone (612-246-3838). For more information about our company,<br />

visit our website at www.regresearch.com.<br />

Is your information up-to-date in the MSA directory? The <strong>Society</strong> is relying more and more on email to<br />

bring you the latest MSA news, awards announcements and other timely information, and our newsletter. To<br />

ensure that you receive <strong>Society</strong> blast emails and the <strong>Inoculum</strong> as soon as it comes out, and so that your colleagues<br />

can keep in touch, please check the accuracy <strong>of</strong> your email address and contact information in the online<br />

directory. This can be accessed via our web site at www.msafungi.org. If you need assistance with updating<br />

your membership information, or help with your membership log-in ID and password, please contact<br />

Kay Rose, Association Manager at Allen Press, at krose@allenpress.com.


MYCOLOGY ON-LINE<br />

Below is an alphabetical list <strong>of</strong> websites featured in <strong>Inoculum</strong>. Those wishing to add sites to this directory or to edit addresses should email<br />

dnatvig@gmail.com. Unless otherwise notified, listings will be automatically deleted after one year (at the editors discretion).<br />

A New Web Page About Tropical Fungi, Hongos Del Parque “El Haya” (58-5)<br />

hongosdelhaya.blogspot.com/<br />

ASCOFrance.com, a very useful site for illustrations <strong>of</strong> ascomycetes including<br />

anamorphs (accessible in both French and English)<br />

asc<strong>of</strong>rance.com/?lang=us<br />

Ascomycota <strong>of</strong> Sweden<br />

www.umu.se/myconet/asco/indexASCO.html<br />

Basidiomycete Research Group (University <strong>of</strong> Helsinki, Finland) studies<br />

systematics, ecology and evolution <strong>of</strong> fungi in forest environment.<br />

www.basidio.fi<br />

Bibliography <strong>of</strong> Systematic Mycology<br />

www.speciesfungorum.org/BSM/bsm.htm<br />

Cold Spring Harbor Laboratory; Meetings & Courses Programs (58-2)<br />

meetings.cshl.edu<br />

Collection <strong>of</strong> 800 Pictures <strong>of</strong> Macro- and Micro-fungi<br />

www.mycolog.com<br />

Cordyceps Website<br />

www.mushtech.org<br />

Cornell Mushroom Blog (58-1)<br />

http://blog.mycology.cornell.edu/<br />

Cortbase (58-2)<br />

andromeda.botany.gu.se/cortbase.html<br />

Corticoid Nomenclatural Database (56-2)<br />

www.phyloinformatics.org/<br />

The Cybertruffle internet server for mycology seeks to provide information<br />

about fungi from a global standpoint (59-3).<br />

www.cybertruffle.org.uk<br />

Cyberliber, a digital library for mycology (59-3).<br />

www.cybertruffle.org.uk/cyberliber<br />

Cybernome provides nomenclatural and taxonomic information about<br />

fungi and their associated organisms, with access to over 548,000 records<br />

<strong>of</strong> scientific names (59-3).<br />

www.cybertruffle.org.uk/cybernome<br />

Dictionary <strong>of</strong> The Fungi Classification<br />

www.indexfungorum.org/names/fundic.asp<br />

Distribution Maps <strong>of</strong> Caribbean Fungi (56-2)<br />

www.biodiversity.ac.psiweb.com/carimaps/index.htm<br />

Entomopathogenic Fungal Culture Collection (EFCC)<br />

www.mushtech.org<br />

Fungal Environmental Sampling and Informatics Network (58-2)<br />

www.bio.utk.edu/fesin/<br />

Fungi <strong>of</strong> Ecuador<br />

www.mycokey.com/Ecuador.html<br />

German <strong>Mycological</strong> <strong>Society</strong> DGfM<br />

www.dgfm-ev.de<br />

Glomeromycota PHYLOGENY<br />

amf-phylogeny.com<br />

MYCO-LICH facilitates mycology and lichenology studies in Iran.<br />

www.myco-lich.com<br />

Mycologia<br />

mycologia.org<br />

Humboldt Institute — Located on the eastern coast <strong>of</strong> Maine, the institute<br />

is known for the extensive series <strong>of</strong> advanced and pr<strong>of</strong>essional-level natural<br />

history seminars it has <strong>of</strong>fered in Maine since 1987, along with ecological<br />

restoration seminars and expeditions to the neotropics. It publishes<br />

the Northeastern Naturalist and Southeastern Naturalist, two scholarly,<br />

peer-reviewed, natural history science journals.<br />

www.eaglehill.us<br />

Website relating to the taxonomy <strong>of</strong> the Hysteriaceae & Mytilinidiaceae<br />

(Pleosporomycetidae, Dothideomycetes, Ascomycota) to facilitate species<br />

identification using a set <strong>of</strong> updated and revised keys based on those first<br />

published by Hans Zogg in 1962. 59(4)<br />

www.eboehm.com/<br />

Index <strong>of</strong> Fungi<br />

www.indexfungorum.org/names/names.asp<br />

Interactive Key to Hypocreales <strong>of</strong> Southeastern United States (57-2)<br />

nt.ars-grin.gov/sbmlweb/fungi/keydata.cfm<br />

ISHAM: the International <strong>Society</strong> for Human and Animal Mycology<br />

www.isham.org<br />

JSTOR (58-3)<br />

jstor.org<br />

Libri Fungorum <strong>Mycological</strong> Publications (58-3)<br />

194.203.77.76/LibriFungorum/<br />

Mold Testing and Identification Services (58-2)<br />

www.pioneer.net/~microbe/abbeylab.html<br />

McCrone Research Institute is an internationally recognized not-for-pr<strong>of</strong>it<br />

institute specializing primarily in teaching applied microscopy. 59(4)<br />

www.mcri.org<br />

Mountain Justice Summer (58-3)<br />

www.MountainJusticeSummer.org<br />

Mycology Education Mart where all relevant mycology courses can be<br />

posted. www2.bio.ku.dk/mycology/courses/<br />

MycoKey<br />

www.mycokey.com<br />

The Myconet Classification <strong>of</strong> the Ascomycota<br />

www.fieldmuseum.org/myconet<br />

New Electronic Journal about mushrooms from Southeast Mexico (61-4)<br />

http://fungavera.blogspot.com<br />

Northeast <strong>Mycological</strong> Federation (NEMF) foray database (58-2)<br />

www.nemfdata.org<br />

Pacific Northwest Fungi — A peer-reviewed online journal for information<br />

on fungal natural history in Alaska, British Columbia, Idaho, Montana,<br />

Oregon and Washington, including taxonomy, nomenclature, ecology,<br />

and biogeography.<br />

www.pnwfungi.org/<br />

Pleurotus spp.<br />

www.oystermushrooms.net<br />

Rare, Endangered or Under-recorded Fungi in Ukraine (56-2)<br />

www.cybertruffle.org.uk/redlists/index.htm<br />

Registry <strong>of</strong> Mushrooms in Art<br />

members.cox.net/mushroomsinart/<br />

Robigalia provides information about field observations, published records<br />

and reference collection specimens <strong>of</strong> fungi and their associated organisms,<br />

with access to over 685,000 records (59-3).<br />

www.cybertruffle.org.uk/robigalia<br />

Searchable database <strong>of</strong> culture collection <strong>of</strong> wood decay fungi (56-6)<br />

www.fpl.fs.fed.us/rwu4501/index.html<br />

Small Things Considered — A microbe blog on microbes in general, but<br />

carries occasional pieces specifically on fungi.<br />

schaechter.asmblog.org/schaechter/<br />

Tree canopy biodiversity project University <strong>of</strong> Central Missouri (58-4)<br />

faculty.cmsu.edu/myxo/<br />

Trichomycete site includes monograph, interactive keys, a complete<br />

database, world literature, etc. (61-4)<br />

www.nhm.ku.edu/~fungi<br />

The TRTC Fungarium (58-1)<br />

bbc.botany.utoronto.ca/ROM/TRTCFungarium/home.php<br />

U.S. National Fungus Collections (BPI)<br />

Complete Mushroom Specimen Database (57-1)<br />

www.ars.usda.gov/ba/psi/sbml<br />

Valhalla provides information about past mycologists, with names, dates <strong>of</strong><br />

birth and death and, in some cases, biographies and/or portraits (59-3).<br />

www.cybertruffle.org.uk/valhalla<br />

Website for the mycological journal Mycena (56-2)<br />

www.mycena.org/index.htm<br />

Wild Mushrooms From Tokyo<br />

www.ne.jp/asahi/mushroom/tokyo/<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 59


CALENDAR OF EVENTS<br />

NOTE TO MEMBERS:<br />

Those wishing to list upcoming mycological courses, workshops, conventions, symposia, and forays<br />

in the Calendar <strong>of</strong> Events should include complete postal/electronic addresses and submit to<br />

<strong>Inoculum</strong> editor Don Natvig at dnatvig@gmail.com.<br />

July 9-13, 2012<br />

The XVII th Biennial Workshop on the Smuts<br />

and Bunts<br />

Shenzhen, Guangdong, China<br />

Contact Dr. Wu Pinshan, smut2012@1<strong>63</strong>.com<br />

July 15-19, 2012<br />

2012 MSA Meeting<br />

Yale University, New Haven, CT<br />

July 31 - Aug 5, 2012<br />

Yeast Genetics and Molecular Biology Meeting<br />

Princeton University, Princeton, NJ<br />

http://www.yeast-meet.org/2012/<br />

August 2-5, 2012<br />

36th Annual Samuel Ristich Foray, NEMF<br />

East Stroudsburg, PA<br />

http://www.nemf.org/files/menu.htm<br />

August 4-8, 2012<br />

<strong>America</strong>n Phytopathological <strong>Society</strong><br />

Providence, RI<br />

http://www.apsnet.org/Pages/default.aspx<br />

60 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

Aug 5-10, 2012<br />

Ecological <strong>Society</strong> <strong>of</strong> <strong>America</strong><br />

Portland, OR<br />

http://www.esa.org/portland/<br />

Sept 3-6, 2012<br />

BMS Fungal Interactions<br />

Alicante, Spain<br />

http://www.britmycolsoc.org.uk/science/scientific-meetings/2012-alicante/<br />

Dec 13-16, 2012<br />

NAMA Foray<br />

Scotts Valley, CA<br />

http://www.namyco.org/<br />

Mar 12-17, 2013<br />

The 27th Fungal Genetics Conference<br />

Asilomar Conference Center<br />

Pacific Grove, CA<br />

http://www.fgsc.net/announc.html<br />

<strong>Mycological</strong> <strong>Society</strong> <strong>of</strong> <strong>America</strong> — Gift Membership Form<br />

Sponsoring a gift membership in MSA <strong>of</strong>fers tangible support both for the re cip i ent <strong>of</strong> the membership as well as for<br />

my col o gy in general. Pro vid ing both Mycologia and <strong>Inoculum</strong>, a gift membership is an excellent way to further the efforts<br />

<strong>of</strong> our mycological col leagues, es pe cial ly those who cannot afford an MSA membership. In addition to a feeling<br />

<strong>of</strong> great sat is fac tion, you also will receive a convenient reminder for renewal <strong>of</strong> the gift membership the following year.<br />

I want to provide an MSA Gift Membership to the following individual:<br />

Name ________________________________________________________________________________________<br />

Institution ______________________________________________________________________________________<br />

Complete Address ______________________________________________________________________________<br />

Phone _____________________ FAX _________________________ Email _______________________<br />

Please send renewal notices to:<br />

(YOUR name) __________________________________________________________________________________<br />

(YOUR address) ________________________________________________________________________________<br />

Phone _______________________ FAX _______________________ Email _______________________<br />

I agree to pay $98* for this membership by check (payable to MSA, drawn on US bank) ___ VISA ___ Mastercard ___<br />

Acct. # _________________ Name (as it appears on card) _____________________________ Exp. date __________<br />

Send this form to: MSA Business Office, PO Box 1897, Lawrence KS 66044<br />

or FAX to (785) 843-1274, Attn: Processing Department<br />

*If this membership is given after June 1, please add $10 to cover postage for past issues.


The <strong>Mycological</strong> <strong>Society</strong> <strong>of</strong> <strong>America</strong><br />

Sustaining Members 2012<br />

The <strong>Society</strong> is extremely grateful for the continuing support <strong>of</strong> its Sus tain ing Members.<br />

Please patronize them and, whenever possible, let their representatives know <strong>of</strong> our appreciation.<br />

Fungi Perfecti<br />

Attn: Paul Stamets<br />

PO Box 7<strong>63</strong>4<br />

Olympia, WA, 98507<br />

(360)426-9292<br />

info@fungi.com<br />

Mycotaxon, Ltd.<br />

Attn: Richard P. Korf<br />

PO Box 264<br />

Ithaca, NY, 14851-0264<br />

(607) 273-0508<br />

info@mycotaxon.com<br />

Triarch, Inc.<br />

Attn: P.L. Conant - President<br />

PO Box 98<br />

Ripon, WI, 54971<br />

(920)748-5125<br />

Sylvan, Inc.<br />

Attn: Mark Wach<br />

Research Dept Library<br />

198 Nolte Drive<br />

Kittanning, PA, 16201<br />

(724)543-3948<br />

mwach@sylvaninc.com<br />

Syngenta Seeds, Inc.<br />

Attn: Rita Kuznia<br />

Dept Head, Plant Pathology<br />

317 330th Street<br />

Stanton, MN, 55018-4308<br />

(507) 6<strong>63</strong>-7<strong>63</strong>1<br />

rita.kuznia@syngenta.com<br />

Genencor Internation, Inc.<br />

Attn: Michael Ward<br />

925 Page Mill Rd<br />

Palo Alto, CA, 94304<br />

(650)846-5850<br />

mward@genencor.com<br />

Novozymes, Inc.<br />

Attn: Wendy Yoder<br />

1445 Drew Ave<br />

Davis, CA, 95618<br />

(530) 757-8110<br />

wty@novozymes.com<br />

BCN Research Laboratories, Inc.<br />

Attn: Emilia Rico<br />

2491 Stock Creek Blvd<br />

Rockford, TN, 37853<br />

(865)558-6819<br />

emirico@msn.com<br />

You are encouraged to inform the Membership Committee (D. Jean Lodge, Chair,<br />

djlodge@caribe.net) <strong>of</strong> firms or foundations that might be approached about Sustaining<br />

Membership in the MSA. Sustaining members have all the rights and privileges<br />

<strong>of</strong> individual members in the MSA and are listed as Sustaining Members in all<br />

issues <strong>of</strong> Mycologia and <strong>Inoculum</strong>.<br />

<strong>Inoculum</strong> <strong>63</strong>(3), June 2012 61


inoculum<br />

The Newsletter<br />

<strong>of</strong> the<br />

<strong>Mycological</strong><br />

<strong>Society</strong> <strong>of</strong> <strong>America</strong><br />

Supplement to Mycologia<br />

Volume <strong>63</strong>, No. 3<br />

June 2012<br />

<strong>Inoculum</strong> is published six times a year in<br />

even numbered months (February, April,<br />

June, August, October, December). Submit<br />

copy to the Editor by email as attachments,<br />

preferably in MS Word. If you submit pictures,<br />

these need to be sent as separate<br />

JPGs or GIFFs, not embedded in the word<br />

document. The Editor reserves the right to<br />

edit copy submitted in accordance with the<br />

policies <strong>of</strong> <strong>Inoculum</strong> and the Council <strong>of</strong> the<br />

<strong>Mycological</strong> <strong>Society</strong> <strong>of</strong> <strong>America</strong>.<br />

Donald O. Natvig, Editor<br />

Department <strong>of</strong> Biology<br />

University <strong>of</strong> New Mexico<br />

Albuquerque, NM 87131<br />

Telephone: (505) 277-5977<br />

Fax: (505) 277-0304<br />

dnatvig@gmail.com<br />

MSA Officers<br />

President, David Hibbett<br />

Department <strong>of</strong> Biology<br />

Clark University<br />

950 Main St.<br />

Worcester, MA 01610<br />

Phone: 508-793-7332<br />

Fax: 508-793-8861<br />

dhibbett@clarku.edu<br />

President Elect, Mary Berbee<br />

Department <strong>of</strong> Botany<br />

University <strong>of</strong> British Columbia<br />

6270 University Blvd.<br />

Vancouver, BC V6T 1Z4, Canada<br />

Phone: 604-822-3780<br />

Fax: 604-822-6089<br />

berbee@interchange.ubc.ca<br />

Vice president, Joey Spatafora<br />

Dept <strong>of</strong> Botany & Plant Pathology<br />

Oregon State University<br />

Corvallis, OR 97331<br />

Phone: 541-737-5304<br />

Fax: 541-737-3573<br />

spatafoj@science.oregonstate.edu<br />

Secretary, Jessie A. Glaeser<br />

USDA-Forest Service<br />

Forest Products Lab<br />

One Gifford Pinchot Dr.<br />

Madison, WI 53726<br />

Phone: 608-231-9215<br />

Fax: 608-231-9592<br />

msasec1@yahoo.com<br />

Treasurer, Marc Cubetta<br />

Department <strong>of</strong> Plant Pathology<br />

Center for Integrated Fungal Research<br />

North Carolina State University<br />

Box 7567 Partners III Room 225<br />

Raleigh, NC 27695<br />

Phone: 919-513-1227<br />

Fax: 919-513-0024<br />

marc_cubeta@ncsu.edu<br />

Past President: Thomas D. Bruns<br />

pogon@berkeley.edu<br />

MSA Homepage: msafungi.org<br />

62 <strong>Inoculum</strong> <strong>63</strong>(3), June 2012<br />

MSA Endowment Funds<br />

Contributions<br />

I wish to contribute $________ to the following named fund(s):<br />

____ Alexopoulos ____ Emerson-Fuller-Whisler ____ Miller<br />

____ Barksdale-Raper ____ Fitzpatrick ____ Thiers<br />

____ Barr ____ Gilbertson ____ Trappe<br />

____ Bigelow ____ Korf ____ Uecker<br />

____ Butler ____ Luttrell ____ Wells<br />

____ Denison<br />

Research Funds<br />

____ Alexander H. and Helen V. Smith Award<br />

____ Myron P. Backus Graduate Award<br />

____ Clark T. Rogerson Award<br />

____ George W. Martin/Gladys E. Baker Award<br />

____ John Rippon Graduate Research Award<br />

____ Undergraduate Research Award<br />

____ Salomón Bartnicki-García Research Award<br />

Other Funds<br />

____ Constantine J. Alexopoulos Prize<br />

____ John S. Karling Lecture Fund<br />

____ Uncommitted Endowment<br />

____ Other (specify)<br />

I wish to pledge $_____________ a year for ____________ years<br />

_____ to the following fund (s) ____________________________<br />

_____ to some other specified purpose ______________________<br />

_____ to the uncommitted endowment<br />

Name: ________________________________________________<br />

Address: _________________________________________________<br />

_________________________________________________<br />

___ Check ____ Credit Card (Visa, MC, etc): ________________<br />

Credit Card No. ____________________ Exp. Date: _________<br />

Signature: __________________________________________<br />

Please send this completed form and your contribution to:<br />

Christy Classi<br />

Association Manager<br />

Allen Press Inc.<br />

810 East 10th St<br />

Lawrence Kansas 66044<br />

Please make checks payable to the<br />

<strong>Mycological</strong> <strong>Society</strong> <strong>of</strong> <strong>America</strong>


An Invitation to Join MSA<br />

THE MYCOLOGICAL SOCIETY OF AMER I CA<br />

(Please print clearly)<br />

2012 MEMBERSHIP FORM<br />

(You may apply for membership on-line at msafungi.org)<br />

Last name ______________________________ First name _________________________________ M.I. ______<br />

Dept./Street _______________________________________________________________________________________<br />

Univ./Organization __________________________________________________________________________________<br />

City __________________________ State/Prov. __________ Country ____________________ ZIP_________________<br />

Telephone: (____)________ ______________ Email _______________________ Fax (____)______________________<br />

TYPE OF MEMBERSHIP<br />

Cyber Memberships<br />

____ Regular $98 (Includes on-line access to Mycologia and <strong>Inoculum</strong>)<br />

____ Student $50 (Includes on-line access to Mycologia and <strong>Inoculum</strong>)<br />

Hardcopy Memberships<br />

____ Regular $98 (Includes print Mycologia, and on-line access<br />

to Mycologia and <strong>Inoculum</strong>)<br />

____ Student $50 (Includes print Mycologia, and on-line access<br />

to Mycologia and <strong>Inoculum</strong>)<br />

____ Sustaining $278 (Includes print Mycologia, and on-line access to Mycologia<br />

and <strong>Inoculum</strong>, plus listing in Mycologia and <strong>Inoculum</strong>)<br />

____ Life $1,500 + $20 for each family member (One-time payment, Includes print<br />

Mycologia, and on-line access to Mycologia and <strong>Inoculum</strong>)<br />

____ Family $98 (Includes one print copy <strong>of</strong> Mycologia, and on-line<br />

access to Mycologia and <strong>Inoculum</strong>)<br />

____ Emeritus $50 (Includes print Mycologia, and on-line access<br />

to Mycologia and <strong>Inoculum</strong>)<br />

Other Memberships<br />

____ Associate $50 (Includes on-line access to <strong>Inoculum</strong>)<br />

____ Emeritus $0 (Includes on-line access to <strong>Inoculum</strong>)<br />

AREAS OF INTEREST<br />

Mark most appropriate area(s)<br />

____ Cell Biology – Physiology (including cytological, ultrastructural, metabolic regulatory and developmental<br />

aspects <strong>of</strong> cells)<br />

____ Ecology – Pathology (including phytopathology, medical mycology, symbiotic associations, saprobic<br />

relationships and community structure/dynamics)<br />

____ Genetics – Molecular Biology (including transmission, population and molecular genetics and molecular<br />

mechanisms <strong>of</strong> gene expression)<br />

____ Systematics – Evolution (including taxonomy, comparative morphology molecular systematics,<br />

phylogenetic inference, and population biology)<br />

PAYMENT<br />

_____ CHECK [Payable to <strong>Mycological</strong> <strong>Society</strong> <strong>of</strong> <strong>America</strong> and<br />

drawn in US dollars on a US bank]<br />

_____ CREDIT CARD: _____ VISA _____ MASTERCARD<br />

Expiration Date: ____________________________________________<br />

Account No: _______________________________________________<br />

Name as it appears on the card: _______________________________<br />

Mail membership form and payment to:<br />

<strong>Mycological</strong> <strong>Society</strong> <strong>of</strong> <strong>America</strong><br />

Attn: Kay Rose<br />

P.O. Box 1897, Lawrence, KS 66044-8897<br />

Phone: (800) 627-0629 or (785) 843-1221<br />

Fax: (800) 627-0326 or (785) 843-1234<br />

Email: krose@allenpress.com

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!