Academia.eduAcademia.edu
Utah State University DigitalCommons@USU Aspen Bibliography Aspen Research 12-31-2020 Fungal Planet Description Sheets: 1112–1181 P. W. Crous Westerdijk Fungal Biodiversity Institute D. A. Cowan University of Pretoria G. Maggs-Kölling Gobabeb-Namib Research Institute N. Yilmaz University of Pretoria E. Larsson University of Gothenburg C. Angelini Herbario Jardín Botánico Nacional See next page additional authors Follow this andfor additional works at: https://digitalcommons.usu.edu/aspen_bib Part of the Agriculture Commons, Ecology and Evolutionary Biology Commons, Forest Sciences Commons, Genetics and Genomics Commons, and the Plant Sciences Commons Recommended Citation Crous, P. W., G. C. Adams, and L. M. Winton. 2020. Fungal Planet description sheets: 1112–1181. Persoonia 45:251-409 [p.324-225]. (New pathogen description affecting aspen in Alaska.) This Article is brought to you for free and open access by the Aspen Research at DigitalCommons@USU. It has been accepted for inclusion in Aspen Bibliography by an authorized administrator of DigitalCommons@USU. For more information, please contact digitalcommons@usu.edu. Authors P. W. Crous, D. A. Cowan, G. Maggs-Kölling, N. Yilmaz, E. Larsson, C. Angelini, T. E. Brandrud, J. D. W. Dearnaley, B. Dima, F. Dovana, and et al. This article is available at DigitalCommons@USU: https://digitalcommons.usu.edu/aspen_bib/7891 Persoonia 45, 2020: 251 – 409 www.ingentaconnect.com/content/nhn/pimj RESEARCH ARTICLE ISSN (Online) 1878-9080 https://doi.org/10.3767/persoonia.2020.45.10 Fungal Planet description sheets: 1112 – 1181 P.W. Crous1,2, D.A. Cowan3, G. Maggs-Kölling4, N. Yilmaz2, E. Larsson5, C. Angelini6, T.E. Brandrud7, J.D.W. Dearnaley8, B. Dima9, F. Dovana10, N. Fechner11, D. García12, J. Gené12, R.E. Halling13, J. Houbraken1, P. Leonard14, J.J. Luangsa-ard15, W. Noisripoom15, A.E. Rea-Ireland16, H. Ševčíková17, C.W. Smyth18, A. Vizzini10, J.D. Adam19, G.C. Adams20, A.V. Alexandrova21,22, A. Alizadeh23, E. Álvarez Duarte24, V. Andjic25, V. Antonín17, F. Arenas26, R. Assabgui27, J. Ballarà28, A. Banwell29, A. BerrafTebbal30, V.K. Bhatt31, G. Bonito32, W. Botha33, T.I. Burgess34, M. Caboň35, J. Calvert36, L.C. Carvalhais36, R. Courtecuisse37, P. Cullington38, N. Davoodian39, C.A. Decock40, R. Dimitrov41, S. Di Piazza42, A. Drenth36, S. Dumez37, A. Eichmeier30, J. Etayo43, I. Fernández44, J.-P. Fiard45, J. Fournier46, S. Fuentes-Aponte47, M.A.T. Ghanbary48, G. Ghorbani49, A. Giraldo50, A.M. Glushakova21,51, D.E. Gouliamova41, J. Guarro12, F. Halleen52, F. Hampe53, M. Hernández-Restrepo1, I. Iturrieta-González12, M. Jeppson5, A.V. Kachalkin21,54, O. Karimi48, A.N. Khalid55, A. Khonsanit15,56, J.I. Kim57, K. Kim47, M. Kiran55, I. Krisai-Greilhuber58, V. Kučera35, I. Kušan59, S.D. Langenhoven60, T. Lebel61, R. Lebeuf62, K. Liimatainen63, C. Linde64, D.L. Lindner65, L. Lombard1, A.E. Mahamedi66, N. Matočec59, A. Maxwell25, T.W. May67, A.R. McTaggart36, M. Meijer1, A. Mešić59, A.J. Mileto19, A.N. Miller68, A. Molia69, S. Mongkolsamrit15, C. Muñoz Cortés24, J. Muñoz-Mohedano26, A. Morte26, O.V. Morozova70, L. Mostert60, R. MostowfizadehGhalamfarsa71, L.G. Nagy72, A. Navarro-Ródenas26, L. Örstadius73, B.E. Overton19, V. Papp74, R. Para75, U. Peintner76, T.H.G. Pham22, A. Pordel77, A. Pošta59, A. Rodríguez26, M. Romberg47, M. Sandoval-Denis1, K.A. Seifert27,78, K.C. Semwal79, B.J. Sewall80, R.G. Shivas36, M. Slovák35,81, K. Smith25, M. Spetik30, C.F.J. Spies82, K. Syme83, K. Tasanathai15,56, R.G. Thorn29, Z. Tkalčec59, M.A. Tomashevskaya54, D. Torres-Garcia12, Z. Ullah55, C.M. Visagie2, A. Voitk84, L.M. Winton85, J.Z. Groenewald1 Key words ITS nrDNA barcodes LSU new taxa systematics Abstract Novel species of fungi described in this study include those from various countries as follows: Australia, Austroboletus asper on soil, Cylindromonium alloxyli on leaves of Alloxylon pinnatum, Davidhawksworthia quintiniae on leaves of Quintinia sieberi, Exophiala prostantherae on leaves of Prostanthera sp., Lactifluus lactiglaucus on soil, Linteromyces quintiniae (incl. Linteromyces gen. nov.) on leaves of Quintinia sieberi, Lophotrichus medusoides from stem tissue of Citrus garrawayi, Mycena pulchra on soil, Neocalonectria tristaniopsidis (incl. Neocalonectria gen. nov.) and Xyladictyochaeta tristaniopsidis on leaves of Tristaniopsis collina, Parasarocladium tasmanniae on leaves of Tasmannia insipida, Phytophthora aquae-cooljarloo from pond water, Serendipita whamiae as endophyte from roots of Eriochilus cucullatus, Veloboletus limbatus (incl. Veloboletus gen. nov.) on soil. Austria, Cortinarius glaucoelotus on soil. Bulgaria, Suhomyces rilaensis from the gut of Bolitophagus interruptus found on a Polyporus sp. Canada, Cantharellus betularum among leaf litter of Betula, Penicillium saanichii from house dust. Chile, Circinella lampensis on soil, Exophiala embothrii from rhizosphere of Embothrium coccineum. China, Colletotrichum cycadis on leaves of Cycas revoluta. Croatia, Phialocephala melitaea on fallen branch of Pinus halepensis. Czech Republic, Geoglossum jirinae on soil, Pyrenochaetopsis rajhradensis from dead wood of Buxus sempervirens. Dominican Republic, Amanita domingensis on litter of deciduous wood, Melanoleuca dominicana on forest litter. France, Crinipellis nigrolamellata (Martinique) on leaves of Pisonia fragrans, Talaromyces pulveris from bore dust of Xestobium rufovillosum infesting floorboards. French Guiana, Hypoxylon hepaticolor on dead corticated branch. Great Britain, Inocybe ionolepis on soil. India, Cortinarius indopurpurascens among leaf litter of Quercus leucotrichophora. Iran, Pseudopyricularia javanii on infected leaves of Cyperus sp., Xenomonodictys iranica (incl. Xenomonodictys gen. nov.) on wood of Fagus orientalis. Italy, Penicillium vallebormidaense from compost. Namibia, Alternaria mirabibensis on plant litter, Curvularia moringae and Moringomyces phantasmae (incl. Moringomyces gen. nov.) on leaves and flowers of Moringa ovalifolia, Gobabebomyces vachelliae (incl. Gobabebomyces gen. nov.) on leaves of Vachellia erioloba, Preussia procaviae on dung of Procavia capensis. Pakistan, Russula shawarensis from soil on forest floor. Russia, Cyberlindnera dauci from Daucus carota. South Africa, Acremonium behniae on leaves of Behnia reticulata, Dothiora aloidendri and Hantamomyces aloidendri (incl. Hantamomyces gen. nov.) on leaves of Aloidendron dichotomum, Endoconidioma euphorbiae on leaves of Euphorbia mauritanica, Eucasphaeria proteae on leaves of Protea neriifolia, Exophiala mali from inner fruit tissue of Malus sp., Graminopassalora geissorhizae on leaves of Geissorhiza splendidissima, Neocamarosporium leipoldtiae on leaves of Leipoldtia schultzii, © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute You are free to share - to copy, distribute and transmit the work, under the following conditions: Attribution: You must attribute the work in the manner specified by the author or licensor (but not in any way that suggests that they endorse you or your use of the work). Non-commercial: You may not use this work for commercial purposes. No derivative works: You may not alter, transform, or build upon this work. For any reuse or distribution, you must make clear to others the license terms of this work, which can be found at http://creativecommons.org/licenses/by-nc-nd/3.0/legalcode. Any of the above conditions can be waived if you get permission from the copyright holder. Nothing in this license impairs or restricts the author’s moral rights. Persoonia – Volume 45, 2020 252 Abstract (cont.) Neocladosporium osteospermi on leaf spots of Osteospermum moniliferum, Neometulocladosporiella seifertii on leaves of Combretum caffrum, Paramyrothecium pituitipietianum on stems of Grielum humifusum, Phytopythium paucipapillatum from roots of Vitis sp., Stemphylium carpobroti and Verrucocladosporium carpobroti on leaves of Carpobrotus quadrifolius, Suttonomyces cephalophylli on leaves of Cephalophyllum pilansii. Sweden, Coprinopsis rubra on cow dung, Elaphomyces nemoreus from deciduous woodlands. Spain, Polyscytalum pini-canariensis on needles of Pinus canariensis, Pseudosubramaniomyces septatus from stream sediment, Tuber lusitanicum on soil under Quercus suber. Thailand, Tolypocladium flavonigrum on Elaphomyces sp. USA, Chaetothyrina spondiadis on fruits of Spondias mombin, Gymnascella minnisii from bat guano, Juncomyces patwiniorum on culms of Juncus effusus, Moelleriella puertoricoensis on scale insect, Neodothiora populina (incl. Neodothiora gen. nov.) on stem cankers of Populus tremuloides, Pseudogymnoascus palmeri from cave sediment. Vietnam, Cyphellophora vietnamensis on leaf litter, Tylopilus subotsuensis on soil in montane evergreen broadleaf forest. Morphological and culture characteristics are supported by DNA barcodes. Article info Received: 15 September 2020; Accepted: 1 October 2020; Published: 19 December 2020. Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands. 2 Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), Faculty of Natural and Agricultural Sciences, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa. 3 Centre for Microbial Ecology and Genomics, Department of Biochemistry, Genetics and Microbiology, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa. 4 Gobabeb-Namib Research Institute, P.O. Box 953, Walvis Bay, Namibia. 5 Biological and Environmental Sciences, and Gothenburg Global Biodiversity Centre, University of Gothenburg, P.O. Box 461, SE-40530 Göteborg, Sweden. 6 Herbario Jardín Botánico Nacional Dr. Rafael Ma. Moscoso, Santo Domingo, Dominican Republic and Via Cappuccini, 78/8 – 33170 Pordenone, Italy. 7 Norwegian Institute for Nature Research, Gaustadalléen 21, NO-0349 Oslo, Norway. 8 Centre for Crop Health, University of Southern Queensland, Toowoomba 4350, Queensland, Australia. 9 Department of Plant Anatomy, Institute of Biology, Eötvös Loránd University, Pázmány Péter sétány 1/C, H-1117, Budapest, Hungary. 10 Department of Life Sciences and Systems Biology, University of Torino, Viale P.A. Mattioli 25, I-10125 Torino, Italy. 11 Queensland Herbarium, Mt Coot-tha Road, Toowong, Brisbane, Queensland 4066, Australia. 12 Mycology Unit, Medical School and IISPV, Universitat Rovira i Virgili (URV), Sant Llorenç 21, 43201 Reus, Tarragona, Spain. 13 Inst. Systematic Botany, New York Botanical Garden, 2900 Southern Blvd, Bronx, NY, 10458-5126 USA. 14 P.O. Box 1193, Buderim 4556 Queensland, Australia. 15 National Center for Genetic Engineering and Biotechnology (BIOTEC), 113 Thailand Science Park, Phahonyothin Rd., Khlong Nueng, Khlong Luang, Pathum Thani 12120, Thailand. 16 University of Tennessee, Knoxville. Knoxville, TN, 37996 USA. 17 Department of Botany, Moravian Museum, Zelný trh 6, 659 37 Brno, Czech Republic. 18 Binghamton University, Binghamton, NY, 13902 USA. 19 205 East Campus Science Center, Lock Haven University, Lock Haven, PA 17745 USA. 20 Department of Plant Pathology, 406D Plant Science Hall, 1875 N. 38th street, University of Nebraska, Lincoln, NE, USA. 21 Lomonosov Moscow State University (MSU), 119234, Leninskie Gory Str. 1/12, Moscow, Russia. 22 Joint Russian-Vietnamese Tropical Research and Technological Center, Hanoi, Vietnam. 23 Department of Plant Protection, Faculty of Agriculture, Azarbaijan Shahid madani University, Tabriz, Iran. 24 Mycology Unit, Microbiology and Mycology Program, Institute of Biomedical Sciences, University of Chile, Santiago, Chile. 25 Department of Agriculture, Water and Environment, 24 Fricker Rd., Perth, 6105 Western Australia, Australia. 26 Departamento de Biología Vegetal (Botánica), Facultad de Biología, Universidad de Murcia, 30100 Murcia, Spain. 27 Biodiversity (Mycology), Agriculture and Agri-Food Canada, Ottawa, ON K1A0C6, Canada. 28 C/ Tossalet de les Forques, 44, E-08600, Berga, Catalonia, Spain. 29 Department of Biology, University of Western Ontario, London, Ontario, N6A 5B7, Canada. 30 MENDELEUM – Institute of Genetics, Mendel University in Brno, Valticka 334, Lednice, 69144, Czech Republic. 31 Navdanya, 105, Rajpur Road, Dehradun, Uttarakhand, India. 1 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 Department of Plant Soil and Microbial Sciences, 1066 Bogue Street, Michigan State University, East Lansing MI, 48824 USA. ARC Plant Health and Protection, Private Bag X134, Queenswood, Pretoria, 0121, South Africa. Phytophthora Science and Management, Centre for Climate Impacted Terrestrial Ecosystems, Harry Butler Institute, Murdoch University, Murdoch, WA 6150, Australia. Institute of Botany, Plant Science and Biodiversity Centre, Slovak Academy of Sciences, Dúbravská cesta 9, SK-84523, Bratislava, Slovakia. Queensland Alliance for Agriculture and Food Innovation (QAAFI), University of Queensland, Ecosciences Precinct, Level 2C East, GPO Box 267, Brisbane 4001, Queensland, Australia. ULR 4515 - LGCgE (Laboratoire de Génie Civil et géo-Environnement), ER4 (Fonctionnement des écosystèmes terrestres anthropisés) - LSVF (Laboratoire des sciences végétales et fongiques), Faculté des sciences pharmaceutiques, Université de Lille, 3, rue du Professeur Laguesse, F-59006 Lille Cedex. The Beeches, Pleck Lane, Kingston Blount, Oxfordshire, OX39 4RU, UK. National Herbarium of Victoria, Royal Botanic Gardens Victoria, South Yarra, Victoria 3141, Australia. Mycothèque de l’Université catholique de Louvain (MUCL, BCCMTM), Earth and Life Institute – ELIM – Mycology, Université catholique de Louvain, Croix du Sud 2 bte L7.05.06, B-1348 Louvain-la-Neuve, Belgium. The Stephan Angeloff Institute of Microbiology, Bulgarian Academy of Sciences, 26 Acad. Georgi Bonchev, Sofia 1113, Bulgaria. University of Genoa, Department of Earth, Environmental and Life Science, Laboratory of Mycology, Corso Europa 26, 16132 Genoa, Italy. Department of Biology, IES Zizur, Ronda S. Cristóbal 196,31180 Zizur Mayor, Navarra, Spain. Myotis-Chile, Duble Almeyda 2010, Ñuñoa, Santiago, Chile. 3/524, résidence les Cyclades, Rue R. Garcin, F-97200 Fort-de-France. Las Muros, 09420 Rimont, France. USDA APHIS PPQ NIS, 10300, Baltimore Avenue, Beltsville, MD 20705, USA. Department of Plant Protection, Faculty of Agronomy, Sari Agricultural Sciences and Natural Resources University, Sari, Iran. Department of Plant Protection, College of Agriculture and Natural Resources, University of Tehran, Karaj 31587-77871, Iran. Radboud University Medical Centre, Geert Grooteplein Zuid 10, 6525 GA Nijmegen, The Netherlands. Mechnikov Research Institute for Vaccines and Sera, 105064, Moscow, Maly Kazenny by-street, 5A, Russia. Plant Protection Division, ARC Infruitec-Nietvoorbij, Private Bag X5026, Stellenbosch, 7599, South Africa. Wetzlarer Strasse 1, 35510 Butzbach, Germany. All-Russian Collection of Microorganisms, G.K. Skryabin Institute of Biochemistry and Physiology of Microorganisms RAS, 142290, Pushchino, pr. Nauki 5, Russia. Department of Botany, University of the Punjab, Quaid-e-Azam campus, Lahore 5090, Pakistan. Plant Microbe Interaction Research Team, Bioscience and Biotechnology for Agriculture, BIOTEC, 113 Thailand Science Park, Pathum Thani 12120, Thailand. Faculty of Computer Science, Dalhousie University, Halifax, Nova Scotia, B3H 4R2, Canada. Department of Botany and Biodiversity Research, University of Vienna, Rennweg 14, 1030 Wien, Austria. Ruđer Bošković Institute, Bijenička cesta 54, HR-10000 Zagreb, Croatia. Department of Pant Pathology, University of Stellenbosch, Private Bag X1, Matieland, 7602, South Africa. Botanic Gardens & State Herbarium, Adelaide, South Australia, Australia. 775, Rang du Rapide Nord, Saint-Casimir, Québec, G0A 3L0, Canada. 253 Fungal Planet description sheets 63 64 65 66 67 68 69 70 71 72 73 74 75 Jodrell Laboratory, Royal Botanic Gardens, Kew, Surrey TW9 3AB, UK. Ecology and Evolution, Research School of Biology, College of Science, The Australian National University, Canberra, ACT, 2601, Australia. One Gifford Pinchot Drive Madison, WI, 53726 USA. Laboratoire de Biologie des Systèmes Microbiens (LBSM), Département des Sciences Naturelles, Ecole Normale Supérieure de Kouba, Alger BP 92, Vieux-Kouba, Alger, Algeria. Royal Botanic Gardens Victoria, Birdwood Ave, Melbourne, VIC 3004, Australia. University of Illinois Urbana-Champaign, Illinois Natural History Survey, 1816 South Oak Street, Champaign, Illinois, 61820, USA. Alette Iversens gate 5, N-3970 Langesund, Norway. Komarov Botanical Institute of the Russian Academy of Sciences, 197376, 2 Prof. Popov Str., Saint Petersburg, Russia. Department of Plant Protection, Shiraz University, Shiraz, Iran Institute of Biochemistry, Biological Research Center, Temesvari krt 62, H-6726 Szeged, Hungary. Lyckans väg 39A, S-29143 Kristianstad, Sweden. Institute of Horticultural Plant Biology, Szent István University, H-1518, Budapest, Hungary. Via Martiri di Via Fani 22, I-61024, Mombaroccio (PU), Italy © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 76 77 78 79 80 81 82 83 84 85 Institute of Microbiology, University of Innsbruck, Technikerstrasse 25, 6020 Innsbruck, Austria. Plant Protection Research Department, Baluchestan Agricultural and Natural Resources Research and Education Centre, AREEO, Iranshahr, Iran. Department of Biology, Carleton University, 1125 Colonel By Drive, Ottawa, ON K1S 5B6, Canada. Department of Biology, College of Sciences, Eritrea Institute of Technology, Mai Nafhi, Asmara, Eritrea. Department of Biology, Temple University, 1900 N. 12th Street, Philadelphia, PA, 19122 USA. Department of Botany, Charles University, Benátská 2, 128 01 Praha, Czech Republic. ARC Plant Health and Protection, Private Bag X5017, Stellenbosch, 7599, South Africa. National Herbarium of Victoria, Royal Botanic Gardens Victoria, South Yarra, Victoria 3141, Australia. 13 Maple St, Humber Village, Newfoundland and Labrador, A2H 2N2, Canada. U.S.D.A. Forest Service, Forest Health Protection, 3700 Airport Way, Fairbanks, AK 99709, USA. 254 Acknowledgements Pedro Crous acknowledges Brett A. Summerell (Royal Botanic Gardens, Sydney, Australia) and Michael J. Wingfield (FABI, University of Pretoria, South Africa), for making several site photographs and field collections available for study. Jan Dijksterhuis is thanked for SEM photomicrographs of Neocalonectria tristaniopsis. Katrina Syme and co-authors thank the curation staff at BRI, MEL, PERTH for their help with loans and processing of collections. Funding for fieldwork and sequencing was provided by the Helen McLellan Fund (RBG Victoria). A. Vizzini thanks R. Berndt (Curator of Fungus Collections, Herbaria Z+ZT) for the loan of specimens. The 2015 collecting trips to Martinique directed by R. Courtecuisse were made possible through financial help from Communauté Territoriale de Martinique, Parc Naturel Régional de Martinique (PNRM) and French national Forestry Office (ONF). The study of Aleksey V. Kachalkin and colleagues was supported by the Russian Science Foundation (grant No. 19-74-10002). Isabel Iturrieta-González and colleagues were partially supported by the Spanish Ministerio de Economía, Industria y Competitividad (grant CGL2017-88094-P). Financial support was provided by the VEGA grant agency (project 2/0061/19) to Viktor Kučera and Marek Slovák. The studies of V. Antonín and H. Ševčíková were enabled by support provided to the Moravian Museum by the Ministry of Culture of the Czech Republic as part of its long-term conceptual development programme for research institutions (MK000094862). Abigail E. Rea-Ireland and colleagues acknowledge the National Fish & Wildlife Foundation, the Pennsylvania Game Commission, Greg Turner for material support on project, Lock Haven University, Temple University, Joseph Calabrese, Jacob Adam, Collin Wesley, Alden Mileto, and Alina Pislar, as well as Karen Hughes for allowing Abigail Rea-Ireland to finish this undergraduate research while starting graduate studies at the University of Tennessee, Knoxville. Jacques Fournier gratefully acknowledges the Parc Naturel Amazonien de Guyane for having initiated, funded and organized the field work in Saül in 2018 and 2019, in the context of the ABC inventory project during which the new species Hypoxylon hepaticolor was collected. Jed Calvert acknowledges support from the Maxim Foundation for travel, collection and help in the discovery of this taxon. The research of Cobus M. Visagie, Rafik Assabgui & Keith A. Seifert was supported by a grant from the Alfred P. Sloan Foundation Program (grant 2014-06-03) on the Microbiology of the Built Environment. Neven Matočec, Ivana Kušan, Ana Pošta, Zdenko Tkalčec, and Armin Mešić are grateful to the Croatian Science Foundation for their financial support under the project grant HRZZ-IP-2018-01-1736 (ForFungiDNA) and to Miro Pucar for his assistance during the fieldwork. Ana Pošta thanks Croatian Science Foundation for their support under the grant HRZZ-2018-09-7081. Shaun D. Langenhoven and colleagues are grateful to the South African Table Grape Industry, Winetech, the National Research Foundation (grant number: 99916) and Technology and Human Resources for Industry Programme (THRIP) for funding. The grant holders acknowledge that opinions, findings and conclusions or recommendations expressed in any publication generated by the NRF-supported research are that of the authors, and that the NRF accepts no liability whatsoever in this regard. The authors would like to thank Meagan van Dyk for help in formatting the article. This study of Daniel Torres-Garcia, Josepa Gené Persoonia – Volume 45, 2020 and Dania García was partially supported by the Spanish Ministerio de Economía, Industria y Competitividad (grant CGL2017-88094-P). Kanoksri Tasanathai and colleagues would like to thank Morakot Tanticharoen and Somvong Tragoonrung, Platform Technology Management Section, National Center for Genetic Engineering and Biotechnology (BIOTEC), Grant No. P19-50231 and CPMO Grant No. P11-00331 for their support of the program Biodiversity studies of entomopathogenic fungi in Thailand. Jean Lodge for her suggestions about the specimens collected in Puerto Rico. The study of Olga V. Morozova was carried out within the framework of a research project of the Komarov Botanical Institute RAS (АААА-А19-119020890079-6) using equipment of its Core Facility Centre ‘Cell and Molecular Technologies in Plant Science’ with the financial support of Russian Foundation for Basic Research (project no. 20-04-00349). The study of Alina V. Alexandrova was supported by Moscow State University Grant for Leading Scientific Schools ‘Depository of the Living Systems’ in the framework of the MSU Development Program. Roy E. Halling acknowledges grants from the National Science Foundation (USA) DEB–0414665, DEB–1020421 and the National Geographic Society Committee for Research and Exploration in grant #8457–08. Logistical support from the Queensland Herbarium (BRI) aided field studies in Queensland. The Queensland Parks and Wildlife Service offered accommodation and orientation on Fraser Island. The staff and resources of the L.B. & D. Cullman Laboratory at the New York Botanical Garden aided in DNA extraction and amplifications. André De Kesel is thanked for providing insights regarding the morphological and developmental terminology of Clémençon (2012). Sandra Abell, Timothy Baroni, Teresa Lebel, Gregory Mueller, Todd Osmundson, and Klaus Querengasser are thanked for field assistance. Sushma Mandava kindly assisted early on in generating tef1 and LSU sequences and Pooja Singh and Olga Khmelnitsky are thanked for assistance in this project with preliminary molecular phylogenetic and morphological assessments. Dilnora Gouliamova and colleagues were supported by a grant from the Bulgarian Science Fund (KP-06-H31/19). The authors express their gratitude for Borislav Guéorguiev from National Museum of Natural History (Sofia, Bulgaria) for the identification of beetles. Asunción Morte is grateful to AEI/FEDER, UE (CGL2016-78946-R) and Fundación Séneca- Agencia de Ciencia y Tecnología de la Región de Murcia (20866/ PI/18) for financial support. Patrick Leonard and John Dearnaley are grateful for the help and advice given by T. Lebel and F. Guard. Milan Spetik and colleagues acknowledge funding by an Internal Grant of Mendel University (IGA-ZF/2020-DP003). The study of Bálint Dima was partly supported by the ELTE Institutional Excellence Program supported by the National Research, Development and Innovation Office (NKFIH-1157-8/2019-DT) in Hungary. Kesiban Özdemir is thanked for sequencing help. Cortinarius glaucoelotus was sequenced within ABOL, subproject HRSFM University of Vienna, supported by the Austrian Federal Ministry of Education, Science and Research. Kamal C. Semwal and Vinod K. Bhatt are grateful to the Uttarakhand State Council for Science and Technology (UCoST), Dehradun, Uttarakhand, India for the financial support provided under project no. UCS&T/R&D/LS-1/1213/4912. The research of Ellen Larsson and Mikael Jeppson was supported by the Swedish Taxonomy Initiative, SLU Artdatabanken (grant 2019.4.3-13). Fungal Planet description sheets 255 0.99 0.97 0.97 0.95 0.91 0.88 0.99 0.99 0.99 0.88 Cortinarius humolens MT773336.1 Cortinarius praetermissus EU684535.1 Cortinarius glaucoelotus sp. nov. - Fungal Planet 1181 Cortinarius elotoides EU056948.1 Cortinarius pseudoglaucopus AY669573.1 Cortinarius austrocyanites KJ635236.1 Cortinarius collocandoides MK277825.1 Cortinarius indopurpurascens sp. nov. - Fungal Planet 1180 Cortinarius porphyropus KC842523.1 Cortinarius kaimanawa KT875193.1 Cortinarius chalybeus KT875194.1 Mycena roseilignicola KP012993.1 Mycena pulchra sp. nov. - Fungal Planet 1163 Inocybe insinuata KY990500.1 Inocybe armeniaca AY380367.1 Inocybe pudica KY990514.1 1 Inocybe ionolepis sp. nov. - Fungal Planet 1158 Inocybe geophylla AY380377.1 Inocybe ionocephala KY990504.1 Inocybe sublilacina KY990520.1 Inocybe lilacina KY990483.1 Inocybe pallidicremea KY990505.1 Parasola schroeterii HQ847114.1 Coprinopsis lagopus FM160730.1 Coprinopsis cineraria KC992963.1 Coprinopsis scobicola HQ847106.1 Coprinopsis bicornis NG_069114.1 Coprinopsis radiata JX118745.1 Coprinopsis macrocephala AY207186.1 Coprinopsis erythrocephala FN396174.1 Coprinopsis rubra sp. nov. - Fungal Planet 1148 Coprinopsis villosa NG_058816.1 Coprinopsis cinerea JQ045877.1 Amanita shennongjiana NG_064596.1 Amanita populiphila NG_057058.1 Amanita pseudovaginata MH486791.1 Amanita betulae MT229878.1 Amanita simulans MN650863.1 Amanita protecta MN820553.1 Amanita retenta NG_064592.1 0.971 Amanita basiana NG_057060.1 0.93 Amanita domingensis sp. nov. - Fungal Planet 1142 Amanita lippiae NG_057062.1 Crinipellis nigricaulis MK277894.1 Crinipellis scabella MK277897.1 Crinipellis pseudosplachnoides MK277895.1 Crinipellis setipes JF930644.1 Crinipellis podocarpi JF930648.1 MT946361 Crinipellis nigrolamellata sp. nov. - Fungal Planet 1149 MT946362 Marasmius palmivorus MN934819.1 Marasmius curreyi FJ917614.1 Marasmius ruforotula FJ917612.1 Hydnaceae Cortinariaceae Mycenaceae Inocybaceae Psathyrellaceae Agaricales 0.99 Cantharellus enelensis KX592712.1 Cantharellus pallens MF797694.1 Cantharellus isabellinus HM750931.1 Cantharellus camphoratus KX592737.1 Cantharellus formosus KX592748.1 Cantharellus spectaculus KM484691.1 Cantharellus lewisii NG_060394.1 Cantharellus betularum sp. nov. - Fungal Planet 1144 Cantharellus amethysteus MN206944.1 Cantharellus ferruginascens KX828815.1 Cantharellales Backusella lamprospora MH866118.1 Amanitaceae Marasmiaceae 0.1 Overview Agaricomycetes phylogeny – part 1 Consensus phylogram (50 % majority rule) of 435 752 trees resulting from a Bayesian analysis of the LSU sequence alignment (130 sequences including outgroup; 979 aligned positions; 571 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. Families and orders are indicated with coloured blocks to the right of the tree. GenBank accession and/or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Backusella lamprospora (GenBank MH866118.1) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 256 0.99 0.1 Overview Agaricomycetes phylogeny (cont.) – part 2 Sebacinaceae Sebacinales Serendipitaceae Russulaceae Russulales 0.86 0.86 0.88 Serendipita indica NG_059912.1 Serendipita whamiae sp. nov. - Fungal Planet 1173 Serendipita vermifera EU625999.1 Craterocolla cerasi KF061265.1 Chaetospermum camelliae EF589737.1 Chaetospermum chaetosporum NG_058876.1 Russula vesca KT933839.1 Russula medullata MT738256.1 Russula medullata MT738257.1 Russula aeruginea MT738258.1 Russula aeruginea MT738259.1 Russula atroaeruginea HKAS53626 Russula ionochlora MT738264.1 Russula grisea MT738262.1 0.97 Russula grisea MT738263.1 Russula atroglauca MT738251.1 Russula atroglauca MT738252.1 Russula atroglauca JMT738248.1 0.95 Russula atroglauca JMT738249.1 Russula atroglauca MT738250.1 LAH36424 LAH35452 0.98 LAH36425 Russula shawarensis sp. nov. - Fungal Planet 1172 LAH36426 LAH35453 Russula galachroa MT738254.1 0.98 Russula galachroa MT738255.1 Russula faustiana MT738253.1 Russula anatina MT738260.1 Russula anatina MT738261.1 Tylopilus alboater AF139708.1 0.99 Tylopilus felleus AF139710.1 1 Tylopilus indecisus AF456820.1 Tylopilus plumbeoviolaceus AF457405.1 Tylopilus subotsuensis sp. nov. - Fungal Planet 1177 Pulchroboletus rubricitrinus MG026638.1 Xerocomoideae sp. MH220316.1 Pulchroboletus sclerotiorum MH257545.1 Aureoboletus thibetanus KJ907381.1 Aureoboletus projectellus AY684158.1 Boletus miniato-olivaceus MH208311.1 Boletus roodyi MH235230.1 REH 9228 Veloboletus limbatus gen. et sp. nov. - Fungal Planet 1178 REH 8746 Phylloporus rubeolus NG_042667.1 Phylloporus rubrosquamosus NG_042668.1 Phylloporus parvisporus JQ967213.1 0.91 Phylloporus orientalis JQ003701.1 Austroboletus mucosus AY612798.1 Austroboletus viscidoviridis KP242282.1 Austroboletus novae-zelandiae KP242256.1 Austroboletus rostrupii KP242258.1 Austroboletus lacunosus KC552057.1 0.99 Austroboletus niveus DQ534622.1 Austroboletus occidentalis KC552059.1 Austroboletus niveus KP242252.1 0.99 Austroboletus dictyotus JX901138.1 Austroboletus aff. mutabilis KF112487.1 0.99 KP242253.1 MT921383.1 KP242247.1 KP242267.1 Austroboletus asper sp. nov. - Fungal Planet 1143 KP242260.1 KC552056.1 KP242246.1 KP242277.1 Boletaceae Boletales 0.98 Persoonia – Volume 45, 2020 257 0.01 Dothideales Cladosporiales Mycosphaerellales Diaporthe perjuncta NG_059064.1 Elsinoe banksiigena NG_064552.1 Elsinoe eelemani KX372296.1 Myriangiales Elsinoe leucopogonis NG_064551.1 Gobabebomyces vachelliae gen. et sp. nov. - Fungal Planet 1113 Lembosiniella eucalyptorum NG_067908.1 Incertae sedis Endosporium aviarium NG_059195.1 Endosporium populi-tremuloides NG_064317.1 Selenophoma australiensis MH874929.1 Aureobasidiaceae 0.99 Cryptocline arctostaphyli MH873458.1 Saccothecium rubi NG_059644.1 0.97 Pseudosydowia eucalypti GQ303327.2 0.97 Pseudosydowia eucalyptorum NG_067893.1 0.91 Moringomyces phantasmae gen. et sp. nov. - Fungal Planet 1140 Kabatiella caulivora MH870057.1 Saccotheciaceae 0.85 Arxiella lunata MH871994.1 Pseudoseptoria donacis MH870852.1 Selenophoma linicola NG_057801.1 Pseudoseptoria donacis MH877798.1 Pseudoseptoria collariana NG_058000.1 ‘Dothiora’ pistaciae NG_057996.1 Endoconidioma rosae-hissaricae NG_059269.1 Endoconidioma populi NG_059198.1 Coniozyma leucospermi EU552113.1 1 Hormonema carpetanum MF611880.1 CPC 38583 0.99 Coniozyma euphorbiae sp. nov. - Fungal Planet 1128 CPC 38551 Neodothiora populina gen. et sp. nov. - Fungal Planet 1141 ‘Dothiora’ mahoniae MH874022.1 ‘Rhizosphaera pini’ EF114708.1 Delphinella strobiligena DQ470977.1 Sydowia polyspora DQ678058.1 0.87 Rhizosphaera pini KY654326.1 Phaeocryptopus nudus EF114700.1 Rhizosphaera macrospora NG_064115.1 Rhizosphaera oudemansii EF114707.1 Plowrightia abietis EF114703.1 Dothioraceae ‘Plowrightia’ periclymeni FJ215702.1 Rhizosphaera kalkhoffii EF114706.1 Xenomeris abietis MH872299.1 Dothiora schizospora MH868980.1 0.97 Dothiora buxi KY511425.1 Dothiora cactacearum KY929176.1 Dothiora pyrenophora KY929179.1 Dothiora phaeosperma KU728550.1 0.89 Dothiora europaea NG_064093.1 Dothiora cytisi NG_059643.1 Dothiora ceratoniae NG_059113.1 Dothiora bupleuricola NG_059133.1 Dothiora oleae MH871116.1 Dothiora agapanthi NG_059132.1 Dothiora infuscans NG_066397.1 Dothiora aloidendri sp. nov. - Fungal Planet 1125 Rachicladosporium luculiae EU040237.1 Rachicladosporium corymbiae NG_067850.1 Cladosporiaceae Rachicladosporium cboliae NG_057851.1 Rachicladosporium inconspicuum NG_059443.1 0.93 Graphiopsis chlorocephala MH874669.1 CPC 38635 0.94 Verrucocladosporium carpobroti sp. nov. - Fungal Planet 1130 CPC 38645 Verrucocladosporium dirinae MH874471.1 Verrucocladosporium visseri NG 068322.1 0.85 Trimmatostroma salinum NG_064176.1 Neocladosporium osteospermi sp. nov. - Fungal Planet 1132 Neocladosporium syringae MT223912.1 Neocladosporium leucadendri NG_057949.1 Davidiellomyces juncicola MN567659.1 Davidiellomyces australiensis NG_059164.1 Phaeothecoidiella illinoisensis NG_069032.1 Phaeothecoidiella missouriensis NG_069924.1 Stomiopeltis syzygii NG_068323.1 Phaeothecoidiellaceae Chaetothyrina artocarpi MF614834.1 Repetophragma zygopetali NG_060158.1 Chaetothyrina guttulata NG_058932.1 0.90 Chaetothyrina spondiadis sp. nov. - Fungal Planet 1145 Chaetothyrina musarum KU710171.1 Myriangiales Fungal Planet description sheets Overview Dothideomycetes phylogeny – part 1 Consensus phylogram (50 % majority rule) of 146 328 trees resulting from a Bayesian analysis of the LSU sequence alignment (233 sequences including outgroup; 824 aligned positions; 341 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. Families and orders are indicated with coloured blocks to the right of the tree. GenBank accession and/or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Diaporthe perjuncta (GenBank NG_059064.1) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 258 0.01 Overview Dothideomycetes phylogeny (cont.) – part 2 Mycosphaerellaceae Massarinaceae Pleosporales 0.97 Juncomyces californiensis MT373351.1 Juncomyces patwiniorum sp. nov. - Fungal Planet 1115 0.95 Ramularia lethalis KX287174.1 Ramularia tovarae KJ504764.1 Ramularia endophylla KF251723.1 Ramularia helminthiae KX287183.1 Ramularia acroptili EU019257.3 Ramularia cerastiicola JF770464.1 Ramularia acris KX287007.1 0.99 Ramularia acroptili GU214689.1 Pachyramichloridium pini NG_057781.1 Chuppomyces handelii GU214437.1 Ruptoseptoria unedonis KF251732.1 Acervuloseptoria ziziphicola NG_057048.1 Ramulariopsis gossypii NG_059692.1 Cercosporella virgaureae KX286979.1 Ramulariopsis pseudoglycines NG_059693.1 Graminopassalora geissorhizae sp. nov. - Fungal Planet 1114 Graminopassalora graminis GQ852621.1 0.88 Neodeightoniella phragmiticola KF777223.1 Cercospora eremochloae HM235406.1 Cercospora ischaemi KM055432.1 0.98 Cercospora capsici MN213758.1 0.94 Cercospora brassicicola MN213759.1 Cercospora tetragoniae MT095119.1 Cercospora punctiformis MH878508.1 Cercospora maculicola KF421117.1 0.86 Cercospora bidentis KF421118.1 Pseudodidymosphaeria spartii KP325437.1 Suttonomyces rosae NG_059882.1 Suttonomyces cephalophylli sp. nov. - Fungal Planet 1127 Stagonospora pseudopaludosa NG_058052.1 Stagonospora tainanensis AB807580.1 1 Stagonospora imperaticola NG_059793.1 Stagonospora pseudoperfecta NG_059399.1 Stagonospora paludosa MH877040.1 Stagonospora trichophoricola MH878585.1 Stagonospora lomandrae NG_058524.1 Stagonospora forlicesenensis NG_059716.1 Stagonospora victoriana NG_058518.1 0.87 Neottiosporina paspali EU754172.1 Stagonospora multiseptata NG_068239.1 Helminthosporium solani MH866589.1 Helminthosporium juglandinum KY984322.1 Helminthosporium genistae KY984316.1 Helminthosporium caespitosum KY984305.1 Helminthosporium oligosporum KY984333.1 Helminthosporium tiliae KY984346.1 Helminthosporium caespitosum KY984306.1 Helminthosporium velutinum KY984355.1 Corynespora leucadendri NG_058860.1 Helminthosporium dalbergiae AB807521.1 Helminthosporium magnisporum AB807522.1 Helminthosporium microsorum KY984329.1 Helminthosporium quercinum KY984338.1 Camarographium carpini NG_058837.1 Preussia dubia GQ203736.1 Sporormia fimetaria GQ203728.1 0.97 Preussia terricola NG_064044.1 Sporormiella megalospora GQ203743.1 Preussia longisporopsis GQ203742.1 Sporormiella pulchella GQ203747.1 Xenomonodictys iranica sp. nov. - Fungal Planet 1179 0.99 Sparticola junci KU721766.1 Neomassarina chromolaenae NG_068715.1 Neomassarina pandanicola NG_069535.1 Neomassarina thailandica MT214467.1 Pleospora iqbalii NG_057738.1 Preussia isomera NG_064045.1 Preussia procaviae sp. nov. - Fungal Planet 1138 Preussia minimoides KF557659.1 Sporormiella isomera MH872355.1 Preussia bipartis GQ203733.1 Preussia australis AB470572.1 0.89 Preussia intermedia MH878451.1 Preussia minipascua GQ203745.1 0.98 Preussia borealis GQ203734.1 Preussia lignicola DQ384098.1 Preussia lignicola MT472604.1 Mycosphaerellales (continued) Persoonia – Volume 45, 2020 Sporormiaceae Fungal Planet description sheets 0.92 0.94 0.94 0.01 Overview Dothideomycetes phylogeny (cont.) – part 3 © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute Neocamarosporiaceae Pyrenochaetopsidaceae Pleosporales (continued) 0.98 Neocamarosporium chichastianum KP004483.1 Neocamarosporium chersinae KY929182.1 Neocamarosporium leipoldtiae sp. nov. - Fungal Planet 1124 Dimorphosporicola tragani KU728536.1 Neocamarosporium salicornicola MF434281.1 Neocamarosporium korfii MF434278.1 Neocamarosporium lamiacearum MF434279.1 Neocamarosporium salsolae MF434282.1 Pyrenochaetopsis indica GQ387626.1 Pyrenochaetopsis botulispora LN907441.1 Pyrenochaetopsis uberiformis LN907420.1 Pyrenochaetopsis poae KJ869175.1 Pyrenochaetopsis decipiens MH872974.1 Pyrenochaetopsis leptospora MH874351.1 Pyrenochaetopsis setosissima GQ387632.1 Pyrenochaetopsis microspora NG_069864.1 Pyrenochaetopsis rajhradensis sp. nov. - Fungal Planet 1171 Alternaria penicillata MH868099.1 Alternaria papavericola MH874555.1 0.91 Septonema secedens MH867983.1 Alternaria alternariae MH875771.1 Alternaria botrytis MK685170.1 Alternaria photistica MH873634.1 Alternaria cinerariae MH878505.1 Alternaria dauci MH873342.1 Alternaria tropica MH874097.1 Alternaria mirabibensis sp. nov. - Fungal Planet 1137 Alternaria helianthiinficiens MH873632.1 Alternaria planifunda MH873356.1 Alternaria hyacinthi KC584332.1 Alternaria proteae MH873911.1 Alternaria zantedeschiae MH875926.1 Alternaria oblongo-obovoidea MH876050.1 Alternaria sorghi MH875927.1 Alternaria cantlous MH874786.1 Alternaria atra MH875550.1 Alternaria terricola NG_069728.1 Alternaria multiformis NG_069860.1 Stemphylium solani MH868918.1 Stemphylium lycopersici MH874764.1 Stemphylium vesicarium MT472605.1 Stemphylium beticola MH878201.1 Stemphylium botryosum NG_069738.1 Stemphylium triglochinicola MH870933.1 Stemphylium paludiscirpi NG_069865.1 Stemphylium eturmiunum NG_069866.1 Stemphylium lycopersici MH874554.1 Stemphylium carpobroti sp. nov. - Fungal Planet 1131 Paradendryphiella salina MH869472.1 Paradendryphiella arenariae MH869281.1 Asteromyces cruciatus MH867592.1 Asteromyces cruciatus KM272363.1 Asteromyces cruciatus NG_064035.1 Bipolaris zeae KY047116.1 Bipolaris drechsleri NG_070031.1 Bipolaris austrostipae NG_069368.1 Bipolaris mediocris MH875914.1 Bipolaris zeicola MN877767.1 Bipolaris multiformis MH875920.1 Bipolaris yamadae MN017990.1 Bipolaris maydis MT516310.1 Bipolaris microlaenae NG_069493.1 Bipolaris setariae MH877691.1 Bipolaris sivanesaniana KX452439.1 Bipolaris cynodontis NG_069234.1 Bipolaris woodii KX452441.1 Bipolaris axonopicola NG_069369.1 Dichotomophthora basellae MK442522.1 Curvularia papendorfii MH875471.1 Dichotomophthora lutea LT990623.1 Dichotomophthora portulacae LT990624.1 Curvularia intermedia HG779038.1 Curvularia tuberculata MH871612.1 0.95 Curvularia oryzae MH871613.1 Curvularia bannonii KJ415497.1 Curvularia pandanicola MH260288.1 Curvularia eragrostidis MH938080.1 Curvularia fallax MH868532.1 259 Pleosporaceae 260 Persoonia – Volume 45, 2020 Chaetothyriales Cyphellophoraceae Herpotrichiellaceae Trichocomaceae Elaphomycetaceae Eurotiales Diaporthe perjuncta NG_059064.1 Cyphellophora reptans NG_067426.1 Cyphellophora sessilis NG_067427.1 Cyphellophora vietnamensis sp. nov. - Fungal Planet 1151 Cyphellophora oxyspora NG_067405.1 Cyphellophora musae NG_068524.1 Cyphellophora pluriseptata NG_067429.1 Cyphellophora guyanensis NG_068580.1 Cyphellophora eucalypti KC455254.1 Cyphellophora phyllostachydis KP122933.1 Cyphellophora europaea KC455259.1 Cyphellophora suttonii KC455256.1 Cyphellophora fusarioides NG_067471.1 0.86 Cyphellophora pauciseptata NG_067428.1 Exophiala brunnea MH870554.1 Veronaea botryosa NG_057788.1 Exophiala lecanii-corni NG_059200.1 Exophiala mali sp. nov. - Fungal Planet 1119 Exophiala mesophila MH877351.1 Exophiala castellanii FJ358241.1 Fonsecaea pedrosoi AF050276.1 Exophiala pisciphila MH872483.1 Exophiala salmonis MH870616.1 Exophiala radicis KT723448.1 Exophiala tremulae JF951155.1 0.94 Exophiala equina MH876297.1 Exophiala prostantherae sp. nov. - Fungal Planet 1122 Exophiala bonariae KR781083.1 0.97 Exophiala cancerae MH874540.1 Exophiala psychrophila MH873750.1 0.91 CBS 146558 CBS 146559 Exophiala embothrii sp. nov. - Fungal Planet 1153 0.95 CBS 146560 CBS 146561 Talaromyces cinnabarinus MH871095.1 Talaromyces purpureogenus MH876655.1 Talaromyces diversus MH876984.1 Talaromyces pulveris sp. nov. - Fungal Planet 1123 Talaromyces rademirici NG_064134.1 Talaromyces purpureus NG_064090.1 Talaromyces pseudostromaticus MH871567.1 Talaromyces pittii MH873420.1 Talaromyces dendriticus MH873068.1 0.92 Talaromyces cecidicola MH874342.1 0.95 Talaromyces coalescens MH876991.1 Elaphomyces papillatus KX238872.1 Elaphomyces granulatus KR029767.1 Elaphomyces asperulus KR029754.1 0.99 Elaphomyces digitatus JN713147.1 Elaphomyces muricatus KR029730.1 Elaphomyces quercicola KX238879.1 0.97 Elaphomyces nemoreus sp. nov. - Fungal Planet 1152 0.97 Elaphomyces decipiens KX238876.1 Penicillium vasconiae NG_064116.1 Penicillium brasilianum NG_069684.1 Penicillium suaveolens MH875702.1 Penicillium sacculum MH878069.1 Penicillium turbatum MH869521.1 Penicillium euglaucum MH872172.1 Penicillium saanichanum sp. nov. - Fungal Planet 1165 Penicillium cinnamopurpureum MH870865.1 Penicillium estinogenum MH876746.1 Penicillium macrosclerotiorum NG_064191.1 0.94 Penicillium vallebormidaense sp. nov. - Fungal Planet 1164 Penicillium citreonigrum MH878369.1 Penicillium erubescens NG_064066.1 Penicillium striatisporum MH872450.1 Penicillium parvum NG_069657.1 Penicillium senticosum MH871920.1 Penicillium parvofructum NG_060189.1 Aspergillaceae 0.01 Overview Eurotiomycetes phylogeny Consensus phylogram (50 % majority rule) of 146 252 trees resulting from a Bayesian analysis of the LSU sequence alignment (70 sequences including outgroup; 838 aligned positions; 267 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. Families and orders are indicated with coloured blocks to the right of the tree. GenBank accession and/or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Diaporthe perjuncta (GenBank NG_059064.1) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). 261 Fungal Planet description sheets Aspergillus niger KC119204.1 Hemileucoglossum littorale KP657566.1 Hemileucoglossum kelabitense MT021912.1 Hemileucoglossum alveolatum KP657565.1 0.99 Hemileucoglossum pusillum MF353093.1 Trichoglossum hirsutum JQ256442.1 Glutinoglossum australasicum KP690100.1 Geoglossum nigritum AY544650.1 Geoglossum fallax JQ256435.1 Geoglossum jirinae sp. nov. - Fungal Planet 1154 0.86 0.99 Geoglossaceae Geoglossales Leucoglossum leucosporum KP272113.1 Geoglossum raitviirii KU986891.1 Geoglossum umbratile JQ256438.1 0.97 0.88 Geoglossum dunense KP744517.1 Geoglossum cookeanum JQ256434.1 Geoglossum heuflerianum KP742955.1 0.99 Geoglossum difforme KC222136.1 Geoglossum simile JQ256437.1 0.01 Overview Geoglossomycetes phylogeny Consensus phylogram (50 % majority rule) of 46 502 trees resulting from a Bayesian analysis of the LSU sequence alignment (18 sequences including outgroup; 930 aligned positions; 223 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. The family and order are indicated with coloured blocks to the right of the tree. GenBank accession or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Aspergillus niger (GenBank KC119204.1) and the taxonomic novelty described in this study for which LSU sequence data were available is indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 262 Persoonia – Volume 45, 2020 0.87 Helotiales Mollisiaceae Dermateaceae Incertae sedis Incertae sedis Thelebolaceae Pseudeurotiaceae Tympanidaceae Thelebolales 0.85 (Rut) / Sclerotiniaceae (Scl) Leotiales 0.96 Xylaria hypoxylon AY544648.1 Lambertella seditiosa KF499359.1 (Rut) Lambertella exophiala KF499363.1 (Rut) 0.87 Staheliella nodosa MH872508.1 (Rut) Lambertella subrenispora AB926152.1 (Rut) Neometulocladosporiella seifertii sp. nov. - Fungal Planet 1129 Lanzia allantospora AB926154.1 (Rut) Neometulocladosporiella eucalypti NG_064541.1 (Rut) Ciboria americana JN086702.1 (Scl) 0.99 Rutstroemiaceae Rutstroemia cuniculi AB926146.1 (Rut) Rutstroemia calopus AB926155.1 (Rut) Rutstroemia paludosa AB926158.1 (Rut) Stromatinia narcissi MH866916.1 (Scl) 0.97 Rutstroemia cunicularia MH878367.1 (Rut) Clarireedia homoeocarpa MH867420.1 (Rut) Stromatinia gladioli MH866477.1 (Scl) Stromatinia gladioli MH866784.1 (Scl) Vibrissea truncorum MT026486.1 Phialocephala fortinii MT026530.1 0.95 Phialocephala cladophialophoroides KY798314.1 Phialocephala dimorphospora MH869755.1 0.99 Phialocephala mallochii MT026545.1 0.94 Phialocephala melitaea sp. nov. - Fungal Planet 1166 0.85 Trimmatostroma salicis EU019300.1 Mollisia cinerea MT026558.1 Mollisia ligni var. ligni MT026520.1 0.92 Obtectodiscus aquaticus MT026501.1 Mollisia endocrystallina NG_068858.1 Mollisia cortegadensis MN129020.1 0.96 Patellariopsis dennisii MK120898.1 Neopyrenopeziza nigripigmentata NG_066459.1 Mollisia cf. cinerea MT026516.1 Acidomelania panicicola NG_064288.1 Neomollisia gelatinosa NG_066452.1 0.95 Mollisia cf. cinerea MH876343.1 Mollisia fusca MH867987.1 Davidhawksworthia quintiniae sp. nov. - Fungal Planet 1116 0.87 Davidhawksworthia ilicicola NG_067307.1 Neofabraea kienholzii KR858873.1 Neofabraea perennans KR858879.1 Pezicula radicicola MH762910.1 Pezicula eucrita KR858968.1 Pezicula carpinea MH867514.1 Pezicula cinnamomea MH877847.1 Pezicula sporulosa KR859049.1 Pezicula neosporulosa MH874192.1 Pezicula cinnamomea MH874303.1 Coleophoma coptospermatis KU728523.1 Coleophoma ericicola KU728528.1 Coleophoma camelliae KU728521.1 Coleophoma coptospermatis KU728522.1 Coleophoma parafusiformis KU728533.1 Coleophoma paracylindrospora KU728532.1 Coleophoma cylindrospora KU728524.1 Coleophoma eucalyptorum KU728529.1 Flagellospora curvula KC834023.1 Bettsia alvei KR139933.1 Ramgea ozimecii NG_068539.1 Thelebolus globosus NG_067263.1 Pseudeurotium hygrophilum JQ780654.1 Pseudeurotium zonatum DQ470988.1 Pseudogymnoascus palmeri sp. nov. - Fungal Planet 1156 Pseudogymnoascus pannorum JQ768405.1 Geomyces auratus NG_042776.1 Tympanis hansbroughiana MH869050.1 Tympanis abietina MK314617.1 Tympanis fasciculata MK314620.1 Tympanis laricina MK314621.1 Tympanis truncatula MK314622.1 Tympanis spermatiospora MK314624.1 Tympanis confusa MK314619.1 Tympanis pitya MK314623.1 Tympanis piceae MH869051.1 0.97 Tympanis tsugae MH869054.1 Tympanis saligna MK314626.1 Tympanis amelanchieris MH869048.1 Tympanis diospyri MH869049.1 Tympanis conspersa MK314618.1 0.01 Overview Leotiomycetes phylogeny Consensus phylogram (50 % majority rule) of 222 002 trees resulting from a Bayesian analysis of the LSU sequence alignment (78 sequences including outgroup; 839 aligned positions; 258 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. Families and orders are indicated with coloured blocks to the right of the tree. Family assignment for Helotiales follows Johnston et al. (2019). GenBank accession or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Xylaria hypoxylon (GenBank AY544648.1) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). 0.99 0.99 Candida broadrunensis KY106372.1 Tuber excavatum DQ191677.1 Tuber neoexcavatum KY013643.1 Tuber buendiae MT102376.1 Tuber melosporum JN392202.1 Tuber pustulatum MK211308.1 0.97 Tuber malacodermum JQ925702.1 Tuber wenchuanense MH115327.1 Tuber huidongense GU979099.1 Tuber furfuraceum GU979090.1 Tuber sinoalbidum MH115299.1 0.88 Tuber microspiculatum MH115316.1 0.93 Tuber umbilicatum GU979086.1 Tuber huidongense GU979097.1 Tuber alcaracense MN953777.1 Tuber aestivum KF523368.1 0.88 Tuber aestivum var. uncinatum MG385627.1 Tuber macrosporum FJ809838.1 Tuber canaliculatum DQ191675.1 Tuber californicum JQ925685.1 Tuber floridanum NG_064427.1 0.99 Tuber lusitanicum sp. nov. - Fungal Planet 1176 Tuber hubeiense NG_059582.1 Tuber pseudomagnatum KP276193.1 Tuber microverrucosum KT067696.1 Tuber anniae JQ925681.1 Tuber liyuanum KT067698.1 Tuber jinshajiangense KX575846.1 Tuber borchii JQ925682.1 Tuber pseudoseparans KT897480.1 0.98 Tuber brunneum KT897478.1 Tuberaceae Pezizales 263 Fungal Planet description sheets 0.01 Overview Pezizomycetes phylogeny Consensus phylogram (50 % majority rule) of 52 502 trees resulting from a Bayesian analysis of the LSU sequence alignment (31 sequences including outgroup; 821 aligned positions; 204 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. The family and order are indicated with coloured blocks to the right of the tree. GenBank accession or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Candida broadrunensis (GenBank KY106372.1) and the taxonomic novelty described in this study for which LSU sequence data were available is indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 264 Persoonia – Volume 45, 2020 Absidia panacisoli NG_063948.1 Phytophthora cinnamomi KX251816.1 Phytophthora sinensis HQ665269.1 Phytophthora melonis KX251712.1 Phytophthora heveae KX251116.1 Phytophthora cryptogea KX251872.1 Phytophthora richardiae KX251928.1 Phytophthora tropicalis HQ665233.1 Phytophthora mengei KX250661.1 0.96 Phytophthora capsici KX250640.1 Phytophthora glovera KX250654.1 Phytophthora mexicana KX250675.1 Phytophthora aysenensis MN557839.1 Phytophthora citricola EU080240.1 0.99 Phytophthora personensis MT159417.1 Phytophthora plurivora KX250836.1 Phytophthora pini KX250815.1 Phytophthora multivora KX250780.1 Phytophthora inflata KX250773.1 Phytophthora citricola KX250752.1 Phytophthora acerina KX250717.1 Phytophthora europaea KX251527.1 Phytophthora alni EU079598.1 Phytophthora rubi KX251569.1 Phytophthora fragariae KX251548.1 Peronosporaceae Phytophthora sulawesiensis EU080349.1 Phytophthora mississippiae KX251310.1 Peronosporales Phytophthora fragariae HQ665150.1 Phytophthora cambivora EU080873.1 Phytophthora bilorbang KX251185.1 Phytophthora gibbosa KX251227.1 Phytophthora gregata KX251255.1 Phytophthora ornamentata KX251324.1 Phytophthora borealis KX251192.1 Phytophthora gonapodyides KX251241.1 0.91 0.98 0.98 Phytophthora riparia KX251352.1 Phytophthora gonapodyides EU080535.1 Phytophthora lacustris KX251276.1 Phytophthora crassamura KX251206.1 0.99 Phytophthora megasperma HQ665228.1 Phytophthora megasperma KX251290.1 Phytophthora rosacearum KX251450.1 0.99 Phytophthora inundata KX251158.1 Phytophthora humicola KX251144.1 0.99 Phytophthora aquae-cooljarloo sp. nov. - Fungal Planet 1167 0.99 Phytophthora gemini KX251130.1 Phytophthora thermophila MG920808.1 Phytophthora amnicola KX251172.1 Phytophthora litoralis KX251283.1 0.97 Phytophthora fluvialis KX251213.1 Phytophthora thermophila KX251359.1 0.01 Overview Phytophthora phylogeny Consensus phylogram (50 % majority rule) of 1 260 002 trees resulting from a Bayesian analysis of the LSU sequence alignment (51 sequences including outgroup; 1 305 aligned positions; 130 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. The family and order are indicated with coloured blocks to the right of the tree. GenBank accession or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Absidia panacisoli (GenBank NG_063948.1) and the taxonomic novelty described in this study for which LSU sequence data were available is indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). 265 Fungal Planet description sheets Backusella lamprospora MH866118.1 Cyberlindnera dauci sp. nov. - Fungal Planet 1150 Cyberlindnera galapagoensis KJ020281.2 Candida mengyuniae EU043158.1 0.95 Cyberlindnera saturnus EF550316.1 Cyberlindnera suaveolens EU544674.1 0.92 Wickerhamomycetaceae Cyberlindnera sargentensis HM461618.1 Cyberlindnera mrakii EF550317.1 Cyberlindnera subsufficiens EF550318.1 Suhomyces taliae KY106790.1 Suhomyces tanzawaensis KY106794.1 Suhomyces atakaporum KY106307.1 Suhomyces panamericanus JQ025407.1 Incertae sedis Suhomyces yuchorum NG_054786.1 Suhomyces chickasaworum NG_054784.1 Suhomyces bolitotheri NG_054783.1 HM627113.1 HM627116.2 Suhomyces bolitophagii sp. nov. - Fungal Planet 1174 HM627061.2 Saccharomycetales Suhomyces terraborum NG_054809.1 0.99 Candida albicans U45776.1 Candida gigantensis AY520316.1 0.87 Candida tropicalis KX198669.1 Candida sojae KJ722420.1 Candida neerlandica NG_054776.1 0.98 0.93 Candida frijolesensis NG_054802.1 Candida labiduridarum NG_042506.1 Debaryomycetaceae Candida tetrigidarum NG_042507.1 Candida prachuapensis NG_054767.1 0.95 0.88 0.93 Candida saraburiensis NG_054769.1 Candida pseudoviswanathii KM586735.1 Candida pellucida MN908679.1 Candida viswanathii U45752.1 0.1 Overview Saccharomycetes phylogeny Consensus phylogram (50 % majority rule) of 69 002 trees resulting from a Bayesian analysis of the LSU sequence alignment (33 sequences including outgroup; 553 aligned positions; 198 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. The families and order are indicated with coloured blocks to the right of the tree. GenBank accession and/or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Backusella lamprospora (GenBank MH866118.1) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 266 0.90 0.95 0.85 0.99 0.97 Ramularia endophylla AY490776.2 Cylindromonium rhabdosporum HQ232120.1 Cylindromonium lichenicola MH871429.1 Phialoseptomonium eucalypti NG_067890.1 Cylindromonium eugeniicola NG_068337.1 Trichonectria setadpressa MT154016.1 Trichonectria rectipila NG_064146.1 Cylindromonium alloxyli sp. nov. - Fungal Planet 1117 Fusarium sarcochroum LT746260.1 Fusarium lateritium var. longum MH866621.1 Fusarium sacchari MH866723.1 Fusarium burgessii MH875064.1 Fusarium oxysporum MH875963.1 Fusarium redolens f. sp. dianthi MH869606.1 Fusarium oxysporum f. sp. gladioli MH866402.1 0.94 Fusarium foetens MH874448.1 Dematiocladium celtidicola NG_059121.1 Gliocephalotrichum cylindrosporum MH871791.1 Neonectria radicicola HQ840378.1 Ilyonectria lusitanica MH876615.1 Cylindrocladiella parva NG_069082.1 Nectriaceae Cylindrocladiella lanceolata MH876849.1 0.99 Cylindrocladiella variabilis MH876847.1 Cylindrocladiella hawaiiensis JN099219.1 0.94 Cylindrocladiella australiensis JN099222.1 Cylindrocladiella viticola JN099226.1 Cylindrocladiella kurandica JN099233.1 Calonectria lauri MH877981.1 Calonectria naviculata MH874551.1 Xenocylindrocladium guianense AY793443.1 Xenocylindrocladium serpens MH876378.1 Calonectria blephiliae KF777197.1 0.99 Calonectria candelabra MH875292.1 Calonectria pentaseptata NG_042733.1 0.94 Calonectria pteridis NG_069022.1 Calonectria amazonica KY653296.1 Curvicladiella cignea JQ666074.1 ‘Curvicladiella’ sp. MT279199.1 CPC 38081 Neocalonectria tristaniopsidis gen. et sp. nov. - Fungal Planet 1112 CPC 38155 Simplicillium calcicola KU746751.1 Simplicillium lamellicola MH866307.1 Simplicillium lanosoniveum MT081959.1 Cordycipitaceae Simplicillium cylindrosporum LC496876.1 Simplicillium obclavatum MH872599.1 Eucasphaeria rustici KY173501.1 Eucasphaeria capensis MH874625.1 Neoeucasphaeria eucalypti NG_067849.1 Niessliaceae Monocillium tenue MG826727.1 Niesslia pulchriseta MG826848.1 Rosasphaeria moravica JF440985.1 Eucasphaeria proteae sp. nov. - Fungal Planet 1133 Smaragdiniseta bisetosa MH873261.1 0.99 Paramyrothecium terrestris KU846333.1 Paramyrothecium foliicola KU846323.1 Paramyrothecium nigrum MH807536.1 Paramyrothecium pituitipietianum sp. nov. - Fungal Planet 1134 Paramyrothecium breviseta KU846319.1 Paramyrothecium roridum MH868182.1 Stachybotryaceae Alfariacladiella spartii KX306777.1 Alfaria caricicola KU845992.1 0.98 Alfaria thymi KU845999.1 Alfaria cyperi-esculenti MG812527.1 0.99 Alfaria acaciae NG_063954.1 Alfaria tabebuiae MK047486.1 Alfaria terrestris KU845996.1 Hypocreales Persoonia – Volume 45, 2020 0.01 Overview Sordariomycetes (Hypocreales) phylogeny – part 1 Consensus phylogram (50 % majority rule) of 1 695 002 trees resulting from a Bayesian analysis of the LSU sequence alignment (135 sequences including outgroup; 812 aligned positions; 341 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. Families and the order are indicated with coloured blocks to the right of the tree. GenBank accession and /or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Ramularia endophylla (GenBank AY490776.2) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). Memnoniella echinata MH870566.1 Memnoniella pseudonilagirica NG_058218.1 Stachybotrys dolichophialis KU846847.1 Stachybotrys limonispora NG_058224.1 Stachybotryaceae (continued) Memnoniella longistipitata KU846182.1 Stachybotrys levispora KP893317.1 0.98 Memnoniella dichroa KU846167.1 Memnoniella oenanthes MH872421.1 Sarocladium bacillisporum MH870718.1 Sarocladium bactrocephalum MH871182.1 1 Sarocladium strictum MH871457.1 Acrodontium salmoneum MH870773.1 Sarocladiaceae 0.96 Parasarocladium radiatum MH869704.1 Parasarocladium breve NG_056979.1 Parasarocladium gamsii NG_056985.1 0.98 Parasarocladium tasmanniae sp. nov. - Fungal Planet 1118 Parasarocladium debruynii NG_066301.1 Acremonium alternatum NG_056977.1 Acremonium behniae sp. nov. - Fungal Planet 1136 Acremonium sclerotigenum MT226553.1 Acremonium charticola MH871527.1 0.91 Acremonium sordidulum NG_056992.1 Acremonium tectonae HQ232144.1 Lasionectria oenanthicola KY607557.1 Acremonium cereale MH877716.1 Lasionectria hilhorstii NG_066302.1 Bionectriaceae Nalanthamala squamicola AF373281.1 Acremonium hyalinulum HQ232045.1 0.91 Hydropisphaera foliicola NG_058273.1 0.92 Paracylindrocarpon pandanicola NG_068837.1 Hydropisphaera erubescens MH876357.1 Paracylindrocarpon aloicola NG_058238.1 Paracylindrocarpon nabanheensis MH376730.1 0.97 Paracylindrocarpon xishuangbannaensis MH376732.1 Tolypocladium sp. KF747144.1 Tolypocladium album MH873866.1 Ophiocordycipitaceae I Tolypocladium capitatum JN941401.1 Claviceps ranunculoides AF245295.1 Moelleriella disjuncta EU392578.1 Moelleriella disjuncta EU392580.1 Moelleriella phyllogena EU392608.1 Moelleriella umbospora EU392628.1 Clavicipitaceae Moelleriella umbospora AY986904.1 0.99 Moelleriella basicystis AY986903.1 0.97 BCC88320 0.89 0.99 Moelleriella puertoricoensis sp. nov. - Fungal Planet 1162 BCC88321 BCC88322 Tolypocladium japonicum DQ518761.1 0.99 MN337287 Tolypocladium flavonigrum sp. nov. - Fungal Planet 1175 MN337288 MN337289 Cordyceps cylindrica EF468841.1 Akanthomyces websteri GQ249979.1 Ophiocordycipitaceae II Hevansia cinerea MH394652.1 Hirsutella citriformis KJ803256.1 Ophiocordyceps entomorrhiza EF468809.1 Hirsutella stilbelliformis var. myrmicarum GQ866967.1 Ophiocordyceps clavata JN941414.1 Hantamomyces aloidendri gen. et sp. nov. - Fungal Planet 1126 Torpedospora radiata AY858948.1 0.99 Glomerulispora mangrovei NG_060628.1 Ophiocordyceps crinalis KF226254.1 0.93 Ophiocordyceps geometridicola MF614647.1 Hirsutella leizhouensis KY415580.1 Ophiocordyceps pauciovoperitheciata MF614649.1 Ophiocordyceps spataforae MG831747.1 Ophiocordyceps coccidiicola AB968419.1 Ophiocordyceps appendiculata JN941412.1 0.99 Ophiocordyceps arborescens NG_060238.1 0.01 Overview Sordariomycetes (Hypocreales) phylogeny (cont.) – part 2 © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 267 Hypocreales (continued) Fungal Planet description sheets 268 Persoonia – Volume 45, 2020 Ramularia endophylla AY490776.2 Macgarvieomyces borealis NG_058088.1 Pyriculariomyces asari MH878225.1 0.99 Pyricularia urashimae NG_059752.1 0.87 Pyricularia caricis MK431456.1 0.86 Pseudopyricularia iraniana NG_060183.1 Dactylaria higginsii DQ341512.1 Pyriculariaceae Pseudopyricularia hagahagae NG_059616.1 Pseudopyricularia javanii sp. nov. - Fungal Planet 1169 Magnaporthales Neopyricularia commelinicola KM484984.1 Pseudopyricularia persiana MH780974.1 Pseudopyricularia hyrcaniana KY457267.1 0.92 Pseudopyricularia bothriochloae NG_058051.1 Colletotrichum colombiense MH876877.1 0.94 Colletotrichum novae-zelandiae MH877051.1 0.98 Colletotrichum fioriniae MH875593.1 Colletotrichum dematium DQ286155.1 0.94 Colletotrichum siamense MK652853.1 Colletotrichum aeschynomenes JX131331.1 Glomerellaceae Colletotrichum lentis NG_069981.1 Colletotrichum gloeosporioides MH876024.1 Glomerellales Colletotrichum spaethianum MH868015.1 Colletotrichum acutatum MH875998.1 371351 0.99 Colletotrichum cycadis sp. nov. - Fungal Planet 1147 371477 Colletotrichum aotearoa MK431454.1 Halosphaeriaceae Lophotrichus medusoides sp. nov. - Fungal Planet 1160 Scedosporium apiospermum FJ345358.1 Petriella setifera DQ470969.1 Microascus trigonosporus KP671717.1 Microascaceae Microascales Magnisphaera stevemossago KT278704.1 Doratomyces stemonitis DQ836907.1 Scopulariopsis brevicaulis KJ443117.1 Harzia cameroonensis KF777216.1 Harzia patula NG_069430.1 Harzia macrospora NG_069429.1 Harzia metrosideri NG_068338.1 1 Ceratostomataceae Melanosporales Harzia verrucosa KY628675.1 Harzia acremonioides NG_067322.1 Harzia tenella NG_066189.1 Harzia sphaerospora NG_067536.1 0.01 Overview Sordariomycetes (Other orders) phylogeny Consensus phylogram (50 % majority rule) of 60 752 trees resulting from a Bayesian analysis of the LSU sequence alignment (41 sequences including outgroup; 807 aligned positions; 247 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. Families and orders are indicated with coloured blocks to the right of the tree. GenBank accession and/or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Ramularia endophylla (GenBank AY490776.2) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). 269 Fungal Planet description sheets Ramularia endophylla AY490776.2 Hypoxylon lenormandii KY610452.1 Hypoxylon fendleri KY610481.1 Hypoxylon hepaticolor sp. nov. - Fungal Planet 1157 Hypoxylaceae Hypoxylon rickii KC968932.1 Daldinia petriniae KY610439.1 0.97 Daldinia vernicosa KY610442.1 Xylaria acuta AY544676.1 0.87 Xylaria hypoxylon AY544648.1 Xylariaceae Halorosellinia oceanica AY083822.1 Rosellinia thelena KF719215.1 Rosellinia aquila KF719207.1 Pseudosubramaniomyces septatus sp. nov. - Fungal Planet 1170 0.88 Pseudosubramaniomyces fusisaprophyticus EU040241.1 Beltraniopsis neolitseae MH878610.1 Subsessila turbinata NG_059724.1 Beltrania rhombica MH869260.1 Porobeltraniella porosa KX519526.1 Beltraniaceae Pseudobeltrania lauri NG_068311.1 Beltraniella endiandrae NG_058665.1 Beltraniella portoricensis MH871777.1 Beltraniella pandanicola MH260281.1 Beltraniella humicola MH870044.1 0.85 Beltraniella podocarpi MT373353.1 Linteromyces quintiniae gen. et sp. nov. - Fungal Planet 1120 Kirstenboschia diospyri NG_057997.1 0.99 0.86 Incertae sedis Xylariales Beltraniella fertilis MF580254.1 Cylindrium grande NG_068656.1 Cylindriaceae Tristratiperidium microsporum NG_070393.1 Cylindrium algarvense MH874925.1 Ciliochorella phanericola NG_067544.1 Sporocadaceae Immersidiscosia eucalypti KY825092.1 Discosia macrozamiae MH327855.1 0.93 Xyladictyochaeta tristaniopsidis sp. nov. - Fungal Planet 1121 Xyladictyochaeta lusitanica MH107972.1 Xyladictyochaetaceae Xyladictyochaeta lusitanica NG_067326.1 Subulispora rectilineata MH872029.1 Phlogicylindrium mokarei NG_059750.1 0.92 Phlogicylindrium dunnii MK442548.1 Phlogicylindrium tereticornis NG_058510.1 Polyscytalum pini-canariensis sp. nov. - Fungal Planet 1135 Polyscytalum neofecundissimum NG_066207.1 Polyscytalum fecundissimum EU035441.1 Phlogicylindriaceae Polyscytalum chilense MH107954.1 Anungitea nullicana NG_057150.1 Phlogicylindrium eucalypti DQ923534.1 Polyscytalum eucalyptigenum NG_057129.1 Anungitea grevilleae KX228304.1 0.94 Anungitea eucalyptorum KJ869176.1 0.01 Overview Sordariomycetes (Xylariales) phylogeny Consensus phylogram (50 % majority rule) of 528 002 trees resulting from a Bayesian analysis of the LSU sequence alignment (49 sequences including outgroup; 815 aligned positions; 218 unique site patterns) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. Families and the order are indicated with coloured blocks to the right of the tree. GenBank accession and/or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Ramularia endophylla (GenBank AY490776.2) and the taxonomic novelties described in this study for which LSU sequence data were available are indicated in bold face. The alignment and tree were deposited in TreeBASE (Submission ID 27179). © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 270 Persoonia – Volume 45, 2020 Neocalonectria tristaniopsidis 271 Fungal Planet description sheets Fungal Planet 1112 – 19 December 2020 Neocalonectria Crous, gen. nov. Etymology. Name refers to its superficial resemblance of the genus Calonectria. Classification — Nectriaceae, Hypocreales, Sordariomycetes. Conidiophores consisting of a stipe, a penicillate arrangement of fertile branches, one to several avesiculate stipe extensions, lacking a terminal vesicle; stipe septate, hyaline, smooth; stipe extensions septate, straight to flexuous, terminating in an aci- cular apical cell. Conidiogenous apparatus: primary branches aseptate or 1-septate, secondary and tertiary branches aseptate, each terminal branch producing 2 – 6 phialides; phialides elongate doliiform to reniform, hyaline, aseptate, apex with minute periclinal thickening and inconspicuous collarette. Conidia cylindrical, rounded at both ends, straight to gently curved, 1-septate, lacking a visible abscission scar, held in parallel cylindrical mucoid clusters. Mega- and microconidia not seen. Type species. Neocalonectria tristaniopsidis Crous. MycoBank MB837819. Neocalonectria tristaniopsidis Crous, sp. nov. Etymology. Name refers to the host genus Tristaniopsis from which it was isolated. Conidiophores consisting of a stipe, a penicillate arrangement of fertile branches, one to several avesiculate stipe extensions, lacking a terminal vesicle; stipe septate, hyaline, smooth, 30–70 × 5–6 µm; stipe extensions septate, straight to flexuous, 70 –150(– 200) µm long, 3 – 4 µm wide at the apical septum, terminating in an acicular apical cell. Conidiogenous apparatus 50 – 80 µm long, 30 – 50 µm wide; primary branches aseptate or 1-septate, 12 – 20 × 4 – 5 µm; secondary branches aseptate, 10 –12 × 3 – 4 µm, and tertiary branches aseptate, 8 –10 × 3–4 µm, each terminal branch producing 2 – 6 phialides; phialides elongate doliiform to reniform, hyaline, aseptate, 8–12 × 2.5–4 µm, apex with minute periclinal thickening and inconspicuous collarette. Conidia cylindrical, rounded at both ends, straight to gently curved, (39 –)40 – 43(– 46) × 3 (– 3.5) µm (mean 42 × 3 µm), 1-septate, lacking a visible abscission scar, held in parallel cylindrical mucoid clusters. Mega- and microconidia not seen. Culture characteristics — Colonies erumpent, spreading, with moderate aerial mycelium and smooth, even margin, covering dish after 2 wk at 25 °C. On MEA, PDA and OA surface ochreous, with chains of brown, thick-walled chlamydospores. Typus. AustrAliA, New South Wales, Limpinwood Nature Reserve, on leaves of Tristaniopsis collina (Myrtaceae), 26 May 2015, B.A. Summerell, HPC 2948 (holotype CBS H-24396, culture ex-type CPC 38081 = CBS 146800, ITS, LSU, actA, cmdA, his3, rpb2, tef1 and tub2 sequences GenBank MW175333.1, MW175373.1, MW173091.1, MW173097.1, MW173106.1, MW173109.1, MW173118.1 and MW173130.1, MycoBank MB837820). Additional material examined. AustrAliA, New South Wales, Limpinwood Nature Reserve, on leaves of T. collina, 26 May 2015, B.A. Summerell, HPC 2948, CBS H-24400, culture CPC 38155 = CBS 146805, ITS, LSU, actA, cmdA, his3, rpb2, tef1 and tub2 sequences GenBank MW175334.1, MW175374.1, MW173092.1, MW173098.1, MW173107.1, MW173110.1, MW173119.1 and MW173131.1. Colour illustrations. Rainforest at Limpinwood Nature Reserve (photo B. Summerell). Penicillate conidiophores giving rise to cylindrical 1-septate conidia on synthetic nutrient-poor agar (scale bars = 10 µm); SEM micrographs captured on host tissue showing conidiophores and conidia (small, aseptate, ellipsoid conidia belong to an acremonium-like fungus). SEM scale bars = 20 µm (left) and 10 µm (right). Notes — Neocalonectria resembles Calonectria and Xenocylindrocladium in having penicillate conidiophores with hyaline, cylindrical, septate conidia (Crous 2002). Morphologically it is closer to Xenocylindrocladium, as it has multiple stipe extensions per conidiophore that lack terminal vesicles (Decock et al. 1997, Crous et al. 2001). Neocalonectria forms a well-supported clade closely related to the genera Calonectria, Curvicladiella and Xenocylindrocladium (Lombard et al. 2015). Although several stipe extensions were observed arising from conidiophores on host material, cultures of Neocalonectria sporulate profusely, but rarely form stipe extensions on synthetic nutrientpoor agar. Morphologically it is hard to argue why the present collection does not belong to the genus Xenocylindrocladium, but phylogenetically, it clusters apart, being more closely related to Curvicladiella, which has hooked, 1-septate, thick-walled, pigmented, verruculose stipe extensions. A Scanning Electron Microscope (SEM) micrograph of Neocalonectria tristaniopsidis can also be seen on the covers of the various issues of Fungal Biology Reviews volume 34, published in 2020. Blast results are supplied as part of the supplementary material. Supplementary material FP1112 Consensus phylogram (50 % majority rule) of 93 002 trees resulting from a Bayesian analysis of the combined 8-gene (ITS, LSU, actA, cmdA, his3, rpb2, tef1 and tub2) sequence alignment (69 sequences including outgroup; 6 214 aligned positions; 418, 203, 347, 600, 395, 668, 504 and 484 unique site patterns, respectively) using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian posterior probabilities (PP) > 0.84 are shown at the nodes and thickened lines represent nodes with PP = 1.00. The scale bar represents the expected changes per site. The taxonomic novelty described in this study is highlighted with bold text and the genera are represented by coloured blocks. The culture collection accession and/or Fungal Planet numbers are indicated behind the species names. The tree was rooted to Stachybotrys chartarum (culture CBS 129.13). The alignment is a reduced version of the alignment used by Lombard et al. (2015) and corresponding GenBank accession numbers of the sequences used can be found in that reference. The alignment and tree were deposited in TreeBASE (Submission ID 27179). Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 272 Persoonia – Volume 45, 2020 Gobabebomyces vachelliae 273 Fungal Planet description sheets Fungal Planet 1113 – 19 December 2020 Gobabebomyces Crous, gen. nov. Etymology. Name refers to the Gobabeb-Namib Research Institute, where this fungus was collected. Classification — Incertae sedis, Myriangiales, Dothideomycetes. Conidiomata erumpent, pycnidial, opening via irregular rupture of epidermis, brown, subglobose, somewhat flattened, exuding a brown conidial mass; wall of 3 – 4 layers of brown textura angularis. Conidiophores reduced to conidiogenous cells lining inner cavity, hyaline, smooth, ampulliform to doliiform, phialidic. Conidia solitary, medium brown, verruculose, aseptate, ellipsoid, thick-walled with obtuse ends. Hyphae hyaline to brown, encased in mucoid sheath, constricted at septa, forming hyaline, smooth, aseptate ellipsoid conidia with obtuse ends, becoming brown and verruculose, and undergoing microcyclic conidiation. Type species. Gobabebomyces vachelliae Crous. MycoBank MB837821. Gobabebomyces vachelliae Crous, sp. nov. Etymology. Name refers to the host genus Vachellia from which it was isolated. Conidiomata restricted to thorns, erumpent, pycnidial, opening via irregular rupture of epidermis, brown, subglobose, somewhat flattened, 80 –150 µm diam, exuding a brown conidial mass; wall of 3 – 4 layers of brown textura angularis. Conidiophores reduced to conidiogenous cells lining inner cavity, hyaline, smooth, ampulliform to doliiform, phialidic, 3 – 5 × 3–4 µm. Conidia solitary, medium brown, verruculose, aseptate, ellipsoid, thick-walled with obtuse ends, (8 –)10 –11(–12) × (5 –)6(–7) µm. In culture hyphae hyaline to brown, 4 – 6 µm diam, encased in mucoid sheath, constricted at septa, forming hyaline, smooth, aseptate ellipsoid conidia with obtuse ends, 5–7 × 3–4 µm, becoming brown and verruculose, swelling and larger in size, and undergoing microcyclic conidiation. Culture characteristics — Colonies erumpent, spreading, surface irregular to folded, with sparse aerial mycelium and uneven margin, reaching 10 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface olivaceous grey, reverse iron-grey. Typus. NAmibiA, Gobabeb-Namib Research Institute, on leaves of Vachellia (= Acacia) erioloba (Fabaceae), 19 Nov. 2019, P.W. Crous, HPC 3132 (holotype CBS H-24450, culture ex-type CPC 38885 = CBS 146779, ITS and LSU sequences GenBank MW175335.1 and MW175375.1, MycoBank MB837822). Notes — Gobabebomyces is an asexual, coniothyrium-like coelomycetous morph related to Lembosiniella, a genus of ascomycetes forming dark brown to black, superficial, irregular leaf spots with linear to Y-shaped hysterothecia on Eucalyptus spp. in Australia (Crous et al. 2019b). Species of Lembosiniella are sterile in culture. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Elsinoe phaseoli (strain CBS 165.31, GenBank MH855166.1; Identities = 388/452 (86 %), 30 gaps (6 %)), Lembosiniella eucalyptorum (strain CBS 144603, GenBank NR_165601.1; Identities = 379/443 (86 %), 24 gaps (5 %)), and Elsinoe australis (strain KNa-5, GenBank FJ010328.2; Identities = 384/451 (85 %), 24 gaps (5 %)). Closest hits using the LSU sequence are Endosporium populi-tremuloides (strain UAMH 10529, GenBank NG_064317.1; Identities = 778/816 (95 %), nine gaps (1 %)), Lembosiniella eucalyptorum (strain CBS 144603, GenBank NG_067908.1; Identities = 774/814 (95 %), six gaps (0 %)), and Elsinoe banksiigena (strain CPC 32402, GenBank NG_064552.1; Identities = 772 /814 (95 %), five gaps (0 %)). Colour illustrations. Vachellia erioloba trees growing at the GobabebNamib Research Institute. Thorn with conidiomata; colonies on malt extract agar; conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl Neriman Yilmaz, Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), Faculty of Natural and Agricultural Sciences, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: neriman.yilmazvisagie@fabi.up.ac.za Don A. Cowan, Centre for Microbial Ecology and Genomics, Department of Biochemistry, Genetics and Microbiology, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: don.cowan@up.ac.za Gillian Maggs-Kölling, Gobabeb-Namib Research Institute, P.O. Box 953, Walvis Bay, Namibia; e-mail: gillian@gobabeb.org © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 274 Persoonia – Volume 45, 2020 Graminopassalora geissorhizae 275 Fungal Planet description sheets Fungal Planet 1114 – 19 December 2020 Graminopassalora geissorhizae Crous, sp. nov. Etymology. Name refers to the host genus Geissorhiza from which it was isolated. Classification — Mycosphaerellaceae, Mycosphaerellales, Dothideomycetes. Sporulating on SNA. Conidiophores medium brown, smooth, fasciculate, arising from a brown stroma of pseudoparenchymatal cells, subcylindrical, branched, 3 –7-septate, up to 160 µm tall, 4–6 µm diam. Conidiogenous cells medium brown, smooth, integrated, subcylindrical, terminal and intercalary, 20–80 × 4–6 µm, with one to several loci, thickened, darkened, refractive, 2 – 3(– 4) µm diam. Conidia solitary, medium brown, smooth to finely verruculose, subcylindrical, straight, apex subobtuse, base truncate, guttulate, 1– 3-septate, (35 –)40 – 55(–70) × (5 –)6 –7 µm; hila thickened, darkened and refractive, (2 –)3 – 4 µm diam. Culture characteristics — Colonies erumpent, spreading, with moderate aerial mycelium and lobed, feathery margin, reaching 15 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse olivaceous grey. Typus. south AfricA, Western Cape Province, Nieuwoudtville, Matjiesfontein, on leaves of Geissorhiza splendidissima (Iridaceae), 2018, P.W. Crous, HPC 3065 (holotype CBS H-24426, culture ex-type CPC 38623 = CBS 146788, ITS, LSU and rpb2 sequences GenBank MW175336.1, MW175376.1 and MW173111.1, MycoBank MB837823). Notes — Graminopassalora, based on G. graminis, is a monotypic genus occurring on members of Poaceae, with conidia 15 – 60 × 5 –14 μm, (0 –)1(– 3)-septate (Braun et al. 2015, Videira et al. 2017). Graminopassalora graminis is widespread on a wide range of grasses, and Deighton (1967) considered G. graminis an aggregate species, possibly composed of several taxa. Graminopassalora geissorhizae is the first member of the genus known from Iridaceae. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Graminopassalora graminis (strain MAFF 510604, GenBank MF951321.1; Identities = 383/411 (93 %), three gaps (0 %)), Pseudocercospora ocimi-basilici (strain ICMP 21324, GenBank MK210535.1; Identities = 377/407 (93 %), two gaps (0 %)), and Pseudocercospora ocimicola (strain CPC 10283, GenBank GU214678.1; Identities = 377/407 (93 %), two gaps (0 %)). Closest hits using the LSU sequence are Graminopassalora graminis (strain CBS 113303, GenBank GU214666.1; Identities = 840/848 (99 %), no gaps), Ramulariopsis pseudoglycines (strain CPC 18242, GenBank NG_059693.1; Identities = 829/848 (98 %), no gaps), and Cercosporella virgaureae (strain CPC 11461, GenBank KX286977.1; Identities = 829/848 (98 %), no gaps). Closest hits using the rpb2 sequence had highest similarity to Zasmidium scaevolicola (strain CBS 127009, GenBank MF951726.1; Identities = 700 /875 (80 %), 25 gaps (2 %)), Zasmidium citri-griseum (strain CBS 122455, GenBank MF951695.1; Identities = 705 /903 (78 %), 26 gaps (2 %)), and Zasmidium hakeicola (strain CBS 144590, GenBank MK442687.1; Identities = 665/860 (77 %), 13 gaps (1 %)). Colour illustrations. Geissorhiza splendidissima with infected leaves. Conidiophores on SNA; conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 276 Persoonia – Volume 45, 2020 Juncomyces patwiniorum 277 Fungal Planet description sheets Fungal Planet 1115 – 19 December 2020 Juncomyces patwiniorum Crous, sp. nov. Etymology. Name refers to the Patwin indigenous people, who are the stewards of the land on which U.C. Davis campus is located. Classification — Mycosphaerellaceae, Mycosphaerellales, Dothideomycetes. Ascomata immersed on culms, globose, brown, 70 –100 µm diam, with central, substomatal ostiole, 15–20 µm diam. Pseudoparaphyses absent. Asci 8-spored, fasciculate, stipitate, fusoid, apex subobtuse, bitunicate, apical chamber absent to 5 µm diam, with basal foot cell present, 75 –100 × 19 – 22 µm. Ascospores multiseriate, fusoid, slightly curved, pale brown, finely verruculose, with large central guttules, constricted at median septum, later becoming 3-septate with obtuse ends, (43 –)50 – 52(– 55) × (5 –)6 µm; germinating from both ends, with germ tubes parallel to the long axis of the spore, not distorting. Asexual morph developing on OA in culture. Mycelium forming chains of subglobose, brown chlamydospores, 8 –12 µm diam, giving rise to erect, unbranched, subcylindrical, straight to slightly curved conidiophores, multiseptate, brown, verruculose, fasciculate, 30–80 × (3–)5–6 µm. Conidiogenous cells terminal, integrated, 15 – 30 × 4 – 6 µm; scars thickened, darkened and refractive, 2–3 µm diam, mostly solitary. Conidia solitary, subcylindrical to narrowly obclavate, slightly flexuous, base truncate, apex subobtuse, medium brown, verruculose, guttulate, (1–)3 – 6-septate, (70 –)80 –120(–130) × (3 –)4 µm; hilum thickened, darkened and refractive, 2.5 – 3 µm diam. Culture characteristics — Colonies erumpent, spreading, surface folded, with moderate aerial mycelium and smooth, lobate margin, reaching 20 mm diam after 2 wk at 25 °C. On MEA surface smoke grey, reverse iron-grey; on PDA surface iron-grey, reverse olivaceous grey; on OA surface olivaceous grey. Notes — Juncomyces represents a monotypic genus in the Mycosphaerellaceae (Videira et al. 2017, Crous et al. 2020b). Juncomyces patwiniorum is characterised by having immersed ascomata with fusoid, slightly curved, pale brown, finely verruculose ascospores that can become 3-septate with age, and a passalora-like asexual morph with pigmented conidia, and darkened, thickened, refractive hila. Juncomyces californiensis is distinct in that it has smaller conidia (45 –)55 –70(–75) × (6 –)7(– 8) μm (Crous et al. 2020b). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Juncomyces californiensis (strain CPC 37989, GenBank MT373368.1; Identities = 482/512 (94 %), 12 gaps (2 %)), Graminopassalora graminis (strain CBS 113303, GenBank GU214666.1; Identities = 455/515 (88 %), 20 gaps (3 %)), and Neokirramyces syzygii (strain CPC 36122, GenBank NR_166317; Identities = 441/496 (89 %), 16 gaps (3 %)). Closest hits using the LSU sequence are Juncomyces californiensis (strain CPC 37989, GenBank MT373351.1; Identities = 808/808 (100 %), no gaps), Xenosonderhenia eucalypti (strain CBS 138858, GenBank MH878634.1; Identities = 865/892 (97 %), three gaps (0 %)), and Ramularia tovarae (strain CBS 113305, GenBank NG_069194.1; Identities = 846/874 (97 %), no gaps). Closest hits using the rpb2 sequence had highest similarity to Ramularia tovarae (strain CBS 113305, GenBank KJ504678.1; Identities = 328/400 (82 %), two gaps (0 %)), Ramularia armoraciae (strain CBS 253.28, GenBank KX288493.1; Identities = 327/401 (82 %), four gaps (0 %)), and Ramularia plurivora (strain CPC 16123, GenBank KJ504653.1; Identities = 325/399 (81 %), no gaps). Typus. usA, California, U.C. Davis campus, on culms of Juncus effusus (Juncaceae), 2 Apr. 2019, P.W. Crous, HPC 2894 (holotype CBS H-24394, culture ex-type CPC 37991 = CBS 146798, ITS, LSU and rpb2 sequences GenBank MW175337.1, MW175377.1 and MW173112.1, MycoBank MB837825). Colour illustrations. Juncus effusus growing on U.C. Davis campus. Immersed ascoma with substomatal ostiole; asci; fusoid, 1-septate ascospores; conidiogenous cells in culture; conidia conidiogenous cells giving rise to conidia. Scale bars: ascoma = 80 µm, all others = 10 µm. Pedro W. Crous, Lorenzo Lombard & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl, l.lombard@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 278 Persoonia – Volume 45, 2020 Davidhawksworthia quintiniae 279 Fungal Planet description sheets Fungal Planet 1116 – 19 December 2020 Davidhawksworthia quintiniae Crous, sp. nov. Etymology. Name refers to the host genus Quintinia from which it was isolated. Classification — Dermateaceae, Helotiales, Leotiomycetes. Mycelium consisting of hyaline, smooth, septate, 2–3 µm diam hyphae. Conidiophores reduced to phialidic conidiogenous cells, ampulliform to doliiform, hyaline, smooth, erect, becoming aggregated in clusters, forming sporodochia on agar surface, 4 –15 × 3 – 5 µm. Conidia solitary, hyaline, smooth, guttulate, aseptate, subcylindrical, ends obtuse, (10 –)11–12(–15) × 2(– 2.5) µm. Culture characteristics — Colonies erumpent, spreading, with sparse aerial mycelium and lobate, smooth margin, reaching 12 mm diam after 2 wk at 25 °C. On MEA surface dirty white, reverse buff; on PDA surface and reverse dirty white; on OA surface buff. Typus. AustrAliA, New South Wales, Limpinwood Nature Reserve, Corina lookout, on leaves of Quintinia sieberi (Paracryphiaceae), 26 May 2015, B.A. Summerell, HPC 2945 (holotype CBS H-24399, culture ex-type CPC 38153 = CBS 146963, ITS, LSU, rpb2 and tub2 sequences GenBank MW175338.1, MW175378.1, MW173113.1 and MW173132.1, MycoBank MB837828). Notes — The erect, ampulliform to doliiform phialides, and hyaline, aseptate conidia are reminiscent of the monotypic genus Davidhawksworthia (Crous & Groenewald 2016). Davidhawksworthia quintiniae is easily distinguished from D. ilicicola (on Ilex aquifolium, Netherlands; conidia 17– 22 × 3 – 3.5 µm) by its smaller conidia. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Dermea libocedri (strain CBS 138.46, GenBank MH856142.1; Identities = 526/560 (94 %), five gaps (0 %)), Dermea acerina (strain CBS 161.38, GenBank MH855942.1; Identities = 524 /560 (94 %), eight gaps (1 %)), Pseudotryblidium neesii (strain HE300, GenBank MK894293.1; Identities = 524/563 (93 %), six gaps (1 %)) and Davidhawksworthia ilicicola (strain CBS 261.95, GenBank KU728516.1; Identities = 517/556 (93 %), 15 gaps (2 %)). Closest hits using the LSU sequence are Davidhawksworthia ilicicola (strain CBS 734.94, GenBank NG_067307.1; Identities = 892/898 (99 %), no gaps), Coleophoma cylindrospora (strain BP-6252, GenBank MH762908.1; Identities = 892/900 (99 %), no gaps), and Coleophoma camelliae (strain CBS 101376, GenBank KU728521.1; Identities = 886/894 (99 %), no gaps). Closest hits using the rpb2 sequence had highest similarity to Rhizodermea veluwensis (strain CBS 110615, GenBank KR859354.1; Identities = 747/882 (85 %), no gaps), Pezicula cornina (strain CBS 285.39, GenBank KR859333.1; Identities = 731/871 (84 %), four gaps (0 %)), and Pezicula neoheterochroma (strain CBS 127388, GenBank KR859338.1; Identities = 744/889 (84 %), no gaps). Closest hits using the tub2 sequence had highest similarity to Davidhawksworthia ilicicola (strain CBS 261.95, GenBank KU728630.1; Identities = 300/371 (81 %), 14 gaps (3 %)), Monilia yunnanensis (strain GND3, GenBank KT736016.1; Identities = 304 /378 (80 %), 17 gaps (4 %)), and Monilinia fructigena (strain CBS 101499, GenBank KT736015.1; Identities = 303/378 (80 %), 17 gaps (4 %)). Colour illustrations. Rainforest at Limpinwood Nature Reserve (photo B. Summerell). Conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 280 Persoonia – Volume 45, 2020 Cylindromonium alloxyli 281 Fungal Planet description sheets Fungal Planet 1117 – 19 December 2020 Cylindromonium alloxyli Crous, sp. nov. Etymology. Name refers to the host genus Alloxylon from which it was isolated. Classification — Nectriaceae, Hypocreales, Sordariomycetes. Mycelium consisting of hyaline, smooth, septate, branched, 1.5–2 µm diam hyphae. Conidiophores solitary, subcylindrical, unbranched, hyaline, smooth, flexuous, erect, 1–2-septate, 30– 60 × 2 µm. Conidiogenous cells integrated, terminal, hyaline, smooth, subcylindrical, 10–20 × 1.5–2 µm, phialidic, apex with periclinal thickening, lacking collarette. Conidia hyaline, smooth, medianly 1-septate, aggregating in cylindrical spore packets, subcylindrical with obtuse ends, (14 –)15 –17(–18) × 2– 3 µm. Culture characteristics — Colonies erumpent, spreading, surface folded, with moderate aerial mycelium and smooth, lobate margin, reaching 25 mm diam after 2 wk at 25 °C. On MEA and PDA surface and reverse saffron; on OA surface buff. Typus. AustrAliA, New South Wales, Limpinwood Nature Reserve, Mt Merino, mycophilic on Meliola on leaves of Alloxylon pinnatum (Proteaceae), 26 May 2015, B.A. Summerell, HPC 2951 (holotype CBS H-24401, culture ex-type CPC 38159 = CBS 146806, ITS, LSU, actA, his3, rpb2, tef1 (first and second part) and tub2 sequences GenBank MW175339.1, MW175379.1, MW173093.1, MW173108.1, MW173114.1, MW173120.1, MW173128.1 and MW173133.1, MycoBank MB837829). Notes — Cylindromonium was recently established as genus to accommodate acremonium-like taxa with unbranched, hyaline conidiophores, and cylindrical, 1-septate conidia (Crous et al. 2019a). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Cylindromonium lichenicola (strain CBS 188.70, GenBank MH859549.1; Identities = 540/591 (91 %), 18 gaps (3 %)), Cylindromonium rhabdosporum (strain CBS 438.66, GenBank MH858850.1; Identities = 538/590 (91 %), 18 gaps (3 %)), and Phialoseptomonium eucalypti (strain CBS 145542, GenBank NR_165572.1; Identities = 534/587 (91 %), 14 gaps (2 %)). Closest hits using the LSU sequence are Cylindromonium lichenicola (strain CBS 415.70A, GenBank MH871536.1; Identities = 588/608 (97 %), two gaps (0 %)), Trichonectria rectipila (strain CBS 132.87, GenBank NG_064146.1; Identities = 583/606 (96 %), two gaps (0 %)), and Phialoseptomonium eucalypti (strain CBS 145542, GenBank NG_067890.1; Identities = 571/596 (96 %), two gaps (0 %)). Closest hits using the tef1 (second part) sequence had highest similarity to Simplicillium aogashimaense (strain JCM 18167, GenBank LC496904.1; Identities = 391/432 (91 %), two gaps (0 %)), Simplicillium cylindrosporum (strain JCM 18169, GenBank LC496906.1; Identities = 390/433 (90 %), two gaps (0 %)), and Nectria marina (strain MFLUCC 16-0544, GenBank MN433214.1; Identities = 389/432 (90 %), no gaps (0 %)).No significant hits were obtained when the actA, his3, rpb2, tef1 (first part) and tub2 sequences were used in blastn and megablast searches. Colour illustrations. Rainforest at Limpinwood Nature Reserve (photo B. Summerell). Conidiophores sporulating on a sterile pine needle; conidiophores and conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 282 Persoonia – Volume 45, 2020 Parasarocladium tasmanniae 283 Fungal Planet description sheets Fungal Planet 1118 – 19 December 2020 Parasarocladium tasmanniae Crous, sp. nov. Etymology. Name refers to the host genus Tasmannia from which it was isolated. Classification — Sarocladiaceae, Hypocreales, Sordariomycetes. Ascomata perithecial, hyaline, smooth-walled, globose to obpyriform, 50 –100 × 50 – 80 µm; wall of 3 – 6 layers of hyaline textura angularis. Asci obovoid to subcylindrical, hyaline, 8-spored, unitunicate with apical mechanism, not straining in Melzer’s reagent, apex slightly flattened, 23 – 30 × 5 – 8 µm, stipitate, intermingled among cellular, hyaline paraphyses that dissolve at maturity. Ascospores hyaline, smooth, guttulate, fusoid-ellipsoid, straight to slightly curved, constricted at median septum, (9 –)10 –11(–12) × (2.5 –)3 (– 3.5) µm. Mycelium consisting of hyaline, smooth, septate, branched, 1.5 – 2 µm diam hyphae. Conidiophores hyaline, smooth, subcylindrical, branched below, 1–2-septate, 20–35 × 2–3 µm. Conidiogenous cells at times solitary, arising directly from superficial hyphae, or on conidiophores, terminal or intercalary, phialidic, subcylindrical with slight apical taper, 12 – 30 × 2 – 3 µm; collarette minute, 1 µm tall, not flared. Conidia hyaline, smooth, granular, subcylindrical to fusoid, straight to slightly curved, aseptate, apex subobtuse, base slightly tapered to truncate hilum, 0.5 µm diam, (5 –)7– 8(– 9) × (1.5–) 2(– 2.5) µm. Culture characteristics — Colonies flat, spreading, surface folded, with sparse aerial mycelium and lobate, even margin, reaching 60 mm diam after 2 wk at 25 °C. On MEA and PDA surface and reverse dirty white; on OA surface buff. Typus. AustrAliA, New South Wales, Limpinwood Nature Reserve, Mt Merino, on leaves of Tasmannia insipida (Winteraceae), 26 May 2015, B.A. Summerell, HPC 2953 (holotype CBS H-24402, culture ex-type CPC 38162 = CBS 146807, ITS, LSU, actA, tef1 and tub2 sequences GenBank MW175340.1, MW175380.1, MW173094.1, MW173121.1 and MW173134.1, MycoBank MB837830). Notes — Parasarocladium was recently introduced by Summerbell et al. (2018) to accommodate a distinct clade of acremonium-like fungi, which are commonly isolated from soil (Crous et al. 2018a). As shown here, however, species can also be foliicolous, and have a sexual morph, which has thus far not been observed for any member of Parasarocladium. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Parasarocladium radiatum (strain CBS 142.62, GenBank NR_161112.1; Identities = 546/581 (94 %), 11 gaps (1 %)), Parasarocladium debruynii (strain CBS 144942, GenBank NR_163316.1; Identities = 540 / 581 (93 %), 17 gaps (2 %)), and Parasarocladium gamsii (as Acremonium gamsii; strain CBS 726.71, GenBank NR_159615.1; Identities = 538/584 (92 %), 17 gaps (2 %)). Closest hits using the LSU sequence are Parasarocladium breve (as Acremonium breve; strain CBS 150.62, GenBank FJ176882.1; Identities = 886/901 (98 %), two gaps (0 %)), Parasarocladium gamsii (strain CBS 726.71, GenBank MH872068.1; Identities = 885/900 (98 %), two gaps (0 %)), and Sarocladium strictum (strain CBS 147.49, GenBank HQ232139.1; Identities = 833 /849 (98 %), four gaps (0 %)). Closest hits using the actA sequence had highest similarity to Parasarocladium debruynii (strain CBS 144942, GenBank MK069413.1; Identities = 594 /648 (92 %), 19 gaps (2 %)), Cordyceps militaris (strain ATCC 34164, GenBank CP023327.1; Identities = 400 / 422 (95 %), no gaps), and Fusarium striatum (strain CBS 101573, GenBank KM231195.1; Identities = 462/515 (90 %), six gaps (1 %)). Closest hits using the tef1 sequence had highest similarity to Parasarocladium debruynii (strain CBS 144942, GenBank MK069410.1; Identities = 241/289 (83 %), 19 gaps (6 %)). Closest hits using the tub2 sequence had highest similarity to Parasarocladium debruynii (strain CBS 144942, GenBank MK069407.1; Identities = 586 /668 (88 %), 27 gaps (4 %)), Sarocladium spirale (strain 3-22, GenBank LC464483.1; Identities = 396 /482 (82 %), 29 gaps (6 %)), and Chaetopsina acutispora (strain CBS 667.92, GenBank KM232029.1; Identities = 317/ 370 (86 %), 16 gaps (4 %)). Colour illustrations. Rainforest at Limpinwood Nature Reserve (photo B. Summerell). Ascomata and ascospores; ascus; conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 284 Persoonia – Volume 45, 2020 Exophiala mali 285 Fungal Planet description sheets Fungal Planet 1119 – 19 December 2020 Exophiala mali Crous, sp. nov. Etymology. Name refers to the host genus Malus from which it was isolated. Classification — Herpotrichiellaceae, Chaetothyriales, Eurotiomycetes. Mycelium consisting of smooth, olivaceous, branched, septate, 2.5–3 µm diam hyphae. Hyphae becoming constricted at septa in terminal region, forming chains of disarticulating conidia, 0 –1-septate, 12 –15 × 3 – 5 µm, subcylindrical to ellipsoid, 0 –1-septate, 8 –10 × 3 – 4 µm, olivaceous, smooth, guttulate. Conidiogenous loci occurring as hyphal pegs on hyphal cells or on conidia, 1– 2 × 1–1.5 µm, not thickened nor darkened, giving rise to smaller, ellipsoid conidia, olivaceous, smooth, aseptate, 4 –7 × 2.5– 3 µm. Culture characteristics — Colonies erumpent, spreading, with moderate aerial mycelium and smooth, lobate margin, reaching 25 mm diam after 2 wk at 25 °C. On MEA surface folded, olivaceous grey, reverse iron-grey; on PDA surface olivaceous grey, reverse iron-grey; on OA surface olivaceous grey. Notes — Species of Exophiala are commonly isolated from soil, water, and plant debris (Crous et al. 2018b). Exophiala mali is a new species of Exophiala that was isolated from apples that underwent cold storage damage due to severe low temperatures. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to ‘Exophiala lecanii-corni’ (strain CMRP3747, GenBank MT452654.1; Identities = 398 / 401 (99 %), no gaps), Exophiala lecanii-corni (strain CBS 123.33, GenBank NR_145351.1; Identities = 560/579 (97 %), five gaps (0 %)), and Exophiala pisciphila (strain G1-2, GenBank KT876529.1; Identities = 592 /610 (97 %), five gaps (0 %)). Closest hits using the LSU sequence are Exophiala lecanii-corni (strain CBS 123.33, GenBank NG_059200.1; Identities = 881/884 (99 %), no gaps), Exophiala pisciphila (strain CBS 464.81, GenBank AF050273.1; Identities = 859 /862 (99 %), no gaps), and Exophiala castellanii (strain CBS 158.58, GenBank NG_070513.1; Identities = 873 /885 (99 %), one gap (0 %)). Typus. south AfricA, Western Cape Province, Ceres, from inner fruit tissue of Malus sp. with cold store damage (Rosaceae), June 2018, P.W. Crous (holotype CBS H-24408, culture ex-type CPC 38208 = CBS 146791, ITS and LSU sequences GenBank MW175341.1 and MW175381.1, MycoBank MB837831). Colour illustrations. Apples with cold store damage. Hyphae and conidiogenous cells; conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 286 Persoonia – Volume 45, 2020 Linteromyces quintiniae 287 Fungal Planet description sheets Fungal Planet 1120 – 19 December 2020 Linteromyces Crous, gen. nov. Etymology. Name refers to the canoe-shaped (L = Linter-) conidia. Classification — Incertae sedis, Xylariales, Sordariomycetes. Mycelium consisting of hyaline, smooth, branched, septate hyphae. Conidiophores reduced to conidiogenous cells or subcylindrical, brown, smooth, erect, unbranched, becoming dark brown and thick-walled with age, septate, with integrated terminal conidiogenous cells. Conidiogenous cells solitary, erect, integrated on hyphae, pale to medium brown, smooth, doliiform to subcylindrical, with several cylindrical denticles near apex. Conidia solitary, aseptate, medium brown, slightly roughened, fusoid, apex and base with apiculus, with paler germ slit along length of conidium body. Type species. Linteromyces quintiniae Crous. MycoBank MB837832. Linteromyces quintiniae Crous, sp. nov. Etymology. Name refers to the host genus Quintinia from which it was isolated. Mycelium consisting of hyaline, smooth, branched, septate, 1.5 – 2 µm diam hyphae. Conidiophores reduced to conidiogenous cells or subcylindrical, brown, smooth, erect, unbranched, becoming dark brown and thick-walled with age, up to 8-septate and 100 µm tall, 3 – 4 µm diam, with integrated terminal conidiogenous cells. Conidiogenous cells solitary, erect, integrated on hyphae, pale to medium brown, smooth, doliiform to subcylindrical, 5 – 20 × 4 – 6 µm, with several cylindrical denticles near apex, 2 – 4 × 1–1.5 µm. Conidia solitary, aseptate, medium brown, slightly roughened, fusoid, apex and base with apiculus, 1– 2 × 1 µm, guttulate, with paler germ slit along length of conidium body, (16 –)20 – 22(– 24) × (6 –)7 µm. Culture characteristics — Colonies erumpent, spreading, surface folded, with moderate aerial mycelium and smooth, lobate margin, reaching 25 mm diam after 2 wk at 25 °C. On MEA surface pale olivaceous grey, reverse olivaceous grey; on PDA surface pale olivaceous grey, reverse olivaceous grey; on OA surface dark brick. Typus. AustrAliA, New South Wales, Limpinwood Nature Reserve, Corina Lookout, on leaves of Quintinia sieberi (Paracryphiaceae), 25 May 2015, B.A. Summerell, HPC 2945 (holotype CBS H-24409, culture ex-type CPC 38231 = CBS 146792, ITS and LSU sequences GenBank MW175342.1 and MW175382.1, MycoBank MB837834). Notes — Linteromyces resembles the genus Subramaniomyces, which has aseptate, polyblastic conidia occurring in branched, acropetal chains on mononematous, branched conidiophores occurring along the length of brown setae. It is morphologically distinct, however, in having solitary conidia, and being phylogenetically unrelated to Subramaniomyces (S. podocarpi, CBS 143176; Crous et al. 2017a), and close to Tristratiperidium, which again has conidia with terminal setulae (Daranagama et al. 2016). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Tristratiperidium microsporum (strain MFLUCC 150413, GenBank NR_164238.1; Identities = 531/581 (91 %), 13 gaps (2 %)), Kiliophora ubiensis (strain IPBCC 131080, GenBank KF056850.1; Identities = 527/579 (91 %), 14 gaps (2 %)), and Kirstenboschia diospyri (strain CBS 134911, GenBank NR_145171.1; Identities = 505/559 (90 %), 17 gaps (3 %)). Closest hits using the LSU sequence are Xyladictyochaeta lusitanica (strain CPC 32526, GenBank MH107973.1; Identities = 818 /844 (97 %), no gaps), Castanediella tereticornis (strain CBS 145068, GenBank NG_068600.1; Identities = 818 /846 (97 %), one gap (0 %)), and Castanediella cagnizarii (strain CBS 101043, GenBank KP858988.1; Identities = 820 /849 (97 %), four gaps (0 %)). Colour illustrations. Rainforest at Limpinwood Nature Reserve (photo B. Summerell). Conidiophores and conidiogenous cells giving rise to conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 288 Persoonia – Volume 45, 2020 Xyladictyochaeta tristaniopsidis 289 Fungal Planet description sheets Fungal Planet 1121 – 19 December 2020 Xyladictyochaeta tristaniopsidis Crous, sp. nov. Etymology. Name refers to the host genus Tristaniopsis from which it was isolated. Classification — Xyladictyochaetaceae, Xylariales, Sordariomycetes. Mycelium consisting of pale brown, smooth, septate, branched, 2 – 3 µm diam hyphae. Conidiophores erect, brown, smooth, subcylindrical, flexuous, multiseptate, 30–100 × 5–6 µm. Conidiogenous cells terminal and intercalary, polyphialidic, 5 –17 × 4–5 µm, phialidic opening 1 µm diam, lacking flared collarettes. Conidia solitary, aggregating in mucoid mass, hyaline, smooth, fusoid-ellipsoid, slightly curved, apex subobtuse, base truncate, 1 µm diam, medianly 1-septate, (16–)17–18(–20) × 2.5(–3) µm; each end with flexuous, unbranched appendage, apex central, base eccentric, 3 – 5 µm diam. Culture characteristics — Colonies flat, spreading, with sparse aerial mycelium and smooth, lobate margin, reaching 40 mm diam after 2 wk at 25 °C. On MEA surface vinaceous buff, reverse isabelline; on PDA surface and reverse dark mouse grey; on OA surface dark mouse grey. Typus. AustrAliA, New South Wales, Limpinwood Nature Reserve, on leaves of Tristaniopsis collina (Myrtaceae), 25 May 2015, B.A. Summerell, HPC 2948 (holotype CBS H-24410, culture ex-type CPC 38240 = CBS 146793, ITS, LSU, tef1 and tub2 sequences GenBank MW175343.1, MW175383.1, MW173122.1 and MW173135.1, MycoBank MB837835). Notes — The monotypic genus Xyladictyochaeta was established by Hernández-Restrepo et al. (2017) to accommodate dictyochaeta-like taxa with terminal and intercalary, polyphialidic conidiogenous cells. Xyladictyochaeta tristaniopsidis has slightly larger conidia than X. lusitanica (11–16 × 2 – 2.5 μm in Hernández-Restrepo et al. (2017); (10 –)11–12(–13) × (2.5 –) 3 μm in Crous et al. (2018b)). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Xyladictyochaeta lusitanica (strain CBS 142290, GenBank NR_154542.1; Identities = 561/575 (98 %), no gaps), Castanediella eucalyptigena (strain CBS 143178, GenBank NR_156384.1; Identities = 536/578 (93 %), 12 gaps (2 %)), and Tristratiperidium microsporum (strain MFLUCC 15-0413, GenBank NR_164238.1; Identities = 527/585 (90 %), nine gaps (1 %)). Closest hits using the LSU sequence are Xyladictyochaeta lusitanica (strain CPC 32526, GenBank MH107973.1; Identities = 784/791 (99 %), no gaps), Castanediella eucalyptigena (strain CBS 143178, GenBank NG_067332.1; Identities = 763/784 (97 %), one gap (0 %)), and Phlogicylindrium eucalypti (strain CBS 120080, GenBank DQ923534.1; Identities = 768/791 (97 %), no gaps). Closest hits using the tef1 sequence had highest similarity to Xyladictyochaeta lusitanica (strain CBS 143502, GenBank MH108033.1; Identities = 467/563 (83 %), 20 gaps (3 %)).Closest hits using the tub2 sequence had highest similarity to Xyladictyochaeta lusitanica (strain CPC 32526, GenBank MH108054.1; Identities = 416/477 (87 %), 15 gaps (3 %)), and Cylindrium aeruginosum (strain CBS 693.83, GenBank KM232124.1; Identities = 295/345 (86 %), 20 gaps (5 %)). Colour illustrations. Rainforest at Limpinwood Nature Reserve (photo B. Summerell). Conidiophores, conidiogenous cells and conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 290 Persoonia – Volume 45, 2020 Exophiala prostantherae Fungal Planet description sheets 291 Fungal Planet 1122 – 19 December 2020 Exophiala prostantherae Crous, sp. nov. Etymology. Name refers to the host genus Prostanthera from which it was isolated. Classification — Herpotrichiellaceae, Chaetothyriales, Eurotiomycetes. Mycelium consisting of pale brown, smooth, branched, septate, 1.5–2 µm diam hyphae. Conidiophores aggregated in clusters, erect, subcylindrical, septate, 5–35 × 2 µm. Conidiogenous cells terminal and intercalary, subcylindrical to cymbiform, phialidic, pale brown, smooth, 4 –12 × 2.5 – 3 µm, apex with minute collarette. Conidia aseptate, guttulate, pale brown, smooth, subcylindrical, apex obtuse, tapering at base to truncate scar, 0.5 µm diam, (3 –)4(– 5) × (1.5–)2 µm. Culture characteristics — Colonies erumpent, spreading, with moderate aerial mycelium and smooth, even margin, reaching 20 mm diam after 2 wk at 25 °C. On MEA and PDA surface olivaceous grey, reverse iron-grey; on OA surface iron-grey. Typus. AustrAliA, New South Wales, Limpinwood Nature Reserve, on leaves of Prostanthera sp. (Lamiaceae), 26 May 2015, B.A. Summerell, HPC 2952 (holotype CBS H-24411, culture ex-type CPC 38251 = CBS 146794, ITS and LSU sequences GenBank MW175344.1 and MW175384.1, MycoBank MB837836). Notes — Exophiala prostantherae is phylogenetically closely related to E. aquamarina (from skin of leafy sea dragon, Phycodures eques, Boston, USA; conidia ellipsoidal to cylindrical, 6.7–19.2 × 4– 4.8 μm; De Hoog et al. 2011) but distinct in having well-defined conidiophores, and smaller conidia. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Exophiala aquamarina (strain IMP-BG-H0001, GenBank MH813288.1; Identities = 550/569 (97 %), four gaps (0 %)), Cadophora fastigiata (strain DN12, GenBank KY781375.1; Identities = 601/633 (95 %), four gaps (0 %)), and Exophiala tremulae (strain CBS 129355, GenBank NR_159874.1; Identities = 600/632 (95 %), four gaps (0 %)). Closest hits using the LSU sequence are Exophiala pisciphila (strain CBS 100.68, GenBank MH870790.1; Identities = 852/856 (99 %), no gaps), Exophiala tremulae (strain UAMH 10998, GenBank JF951155.1; Identities = 852/856 (99 %), no gaps), and Exophiala equina (strain CBS 128222, GenBank MH876297.1; Identities = 851/856 (99 %), no gaps). Colour illustrations. Rainforest at Limpinwood Nature Reserve (photo B. Summerell). Conidiophores, conidiogenous cells and conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 292 Persoonia – Volume 45, 2020 Talaromyces pulveris Fungal Planet description sheets 293 Fungal Planet 1123 – 19 December 2020 Talaromyces pulveris Crous, sp. nov. Etymology. Name refers to the bore dust (L = pulvis) of a beetle, from which it was isolated. Classification — Trichocomaceae, Eurotiales, Eurotiomycetes. Mycelium consisting of hyaline, smooth, branched, septate, 1.5 – 2 µm diam hyphae. Conidiophores dimorphic. Microconidiophores monoverticillate or as solitary phialides, short, 1– 2septate, arising from superficial hyphae, 15–25 × 2 µm. Macroconidiophores biverticillate, erect, subcylindrical, flexuous, penicillate, 70–300 × 2.5–3 µm, stipe smooth to slightly roughened, multiseptate. Metulae two to seven, subcylindrical, hyaline, smooth to slightly roughened, subcylindrical, aseptate, 8 –12 × 2 – 3 µm; additional branches rarely observed. Conidiogenous cells phialidic, arranged in whorls of 2–6 per metula, acerose to subcylindrical with apical taper in upper third, 8 –13 × 2 – 3 µm. Conidia arranged in long, unbranched chains, aseptate, green in masse, in basipetal chains, subglobose, thick-walled, smooth, 2 – 2.5 µm diam. Culture characteristics — Colony diam, 7 d, in mm: CYA, 25 °C: Colonies restricted, non-sulcate, flat, thin; margin entire; mycelium white; sporulation absent or very sparsely produced; soluble pigments absent; exudates absent; reverse white. YES, 25 °C: Similar to CYA, though sporulation lacking. MEA, 25 °C: Colonies non-sulcate, moderately high; margin entire; mycelium white; sporulation sparse to strong; texture floccose to slightly funiculose; soluble pigments absent after 7 d, present after 2 wk, red; exudates absent; conidial colour en masse greygreen; reverse brown or dark brown. DG18, 25 °C: See CYA. OA, 25 °C: Similar to CYA, though sporulation poor to moderate; red soluble pigments produced after 2 wk. Colony diam, 7 d, in mm – CYA microcolonies to 5; CYA30°C microcolonies – 5; CYA37°C no growth; CYAS no growth; DG18 3 – 6; MEA 7–10; OA 5– 8; YES no growth or microcolonies; CREA no growth. Typus. frANce, from bore dust of deathwatch beetle (Xestobium rufovillosum) infesting floorboards (Quercus wood), 2019, C.A. Decock (holotype CBS H-24417, culture ex-type CPC 38523 = MUCL pd8781 = DTO 432H1 = CBS 146831, ITS, LSU, cmdA, rpb2 and tub2 sequences GenBank MW175345.1, MW175385.1, MW173099.1, MW173115.1 and MW173136.1, MycoBank MB837837). Notes — Talaromyces pulveris represents a new species in section Purpurei, phylogenetically most closely related to T. iowaense (Samson et al. 2011, Yilmaz et al. 2014, Crous et al. 2018a, Guevara-Suarez et al. 2020). Talaromyces rademirici is a sister species of T. pulveris and T. iowaense (Samson et al. 2011, Yilmaz et al. 2014, Crous et al. 2018a, Guevara-Suarez et al. 2020, Houbraken et al. 2020). Talaromyces iowaense, T. pulveris and T. rademirici grow restrictedly on CYA and are unable to grow on CYA supplemented with 5 % NaCl. Talaromyces pulveris grows more restricted on MEA (7–10 mm) than T. iowaense (17–18 mm) and T. rademirici (14 –15 mm). The production of subglobose conidia and inability of T. pulveris to grow on CYA incubated at 37 °C is shared with T. iowaense. In contrast, the conidia of T. rademirici are ellipsoidal and this species is able to grow on CYA incubated at 37 °C. Talaromyces iowaense is grows on CREA, while T. pulveris and T. rademirici are unable to grow on this medium (Yilmaz et al. 2014, Crous et al. 2018a). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Talaromyces pseudostromaticus (strain AS3.16005, GenBank MT182956.1; Identities = 472/540 (87 %), 33 gaps (6 %)), Talaromyces pittii (strain CBS 139.84, GenBank MH861710.1; Identities = 472 /540 (87 %), 33 gaps (6 %)), and Talaromyces aculeatus (strain BCC<THA> 88118, GenBank MH997879.1; Identities = 465 /532 (87 %), 30 gaps (5 %)). Closest hits using the LSU sequence are Talaromyces purpureus (strain CBS 475.71, GenBank NG_064090.1; Identities = 787/ 807 (98 %), one gap (0 %)), Talaromyces rademirici (strain CBS 140.84, GenBank NG_064134.1; Identities = 815 /837 (97 %), one gap (0 %)), and Talaromyces dendriticus (strain CBS 660.80, GenBank MH873068.1; Identities = 781/ 806 (97 %), one gap (0 %)). Closest hits using the cmdA sequence had highest similarity to Talaromyces purpureus (strain CBS 475.71, GenBank KJ885292.1; Identities = 354/410 (86 %), 16 gaps (3 %)), Talaromyces ptychoconidium (strain CV2807, GenBank JX140699.1; Identities = 443/550 (81 %), 50 gaps (9 %)), and Talaromyces cecidicola (strain CBS 101419, GenBank KJ885287.1; Identities = 285/ 336 (85 %), seven gaps (2 %)). Closest hits using the rpb2 sequence had highest similarity to Talaromyces rademirici (strain CBS 140.84, GenBank KM023302.1; Identities = 697/760 (92 %), no gaps), Talaromyces purpureus (strain CBS 475.71, GenBank JN121522.1; Identities = 748/825 (91 %), no gaps), and Talaromyces ptychoconidium (strain DTO180F1, GenBank MK450880.1; Identities = 772 /864 (89 %), no gaps. Closest hits using the tub2 sequence had highest similarity to Talaromyces rademirici (strain CBS 140.84, GenBank KJ865734.1; Identities = 395/444 (89 %), nine gaps (2 %)), Talaromyces iowaense (as Talaromyces sp. GP-2018a; strain EMSL 2233, GenBank MH282578.1; Identities = 390/452 (86 %), 12 gaps (2 %)), and Talaromyces ptychoconidium (as Penicillium sp. CMV-2008c; strain CV323, GenBank GU385735.1; Identities = 297/ 346 (86 %), 14 gaps (4 %)). Colour illustrations. Sampling site in France. Colony on MEA; conidiophores and conidiogenous cells giving rise to conidial chains. Scale bars = 10 µm. Pedro W. Crous, Jos Houbraken, M. Meijer & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl, j.houbraken@wi.knaw.nl, m.meijer@wi.knaw.nl & e.groenewald@wi.knaw.nl Cony A. Decock, Mycothèque de l’Université catholique de Louvain (MUCL, BCCMTM), Earth and Life Institute – ELIM – Mycology, Université catholique de Louvain, Croix du Sud 2 bte L7.05.06, B-1348 Louvain-la-Neuve, Belgium; e-mail: cony.decock@uclouvain.be © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 294 Persoonia – Volume 45, 2020 Neocamarosporium leipoldtiae Fungal Planet description sheets 295 Fungal Planet 1124 – 19 December 2020 Neocamarosporium leipoldtiae Crous, sp. nov. Etymology. Name refers to the host genus Leipoldtia from which it was isolated. Classification — Neocamarosporiaceae, Pleosporales, Dothideomycetes. Conidiomata solitary, erumpent, globose, 200 – 300 µm diam, with central ostiole; wall covered by brown, verruculose hyphae, 3 – 4 µm diam; wall consisting of 6 – 8 layers of brown textura angularis. Conidiophores reduced to conidiogenous cells or with a supporting cell, lining the inner cavity, hyaline, smooth, ampulliform with long cylindrical apical part, proliferating percurrently near apex, 12 – 35 × 5 –7 µm. Conidia solitary, medium brown, ellipsoid to subcylindrical, apex obtuse, base truncate, muriformly septate, with 3 – 6 transverse septa, 2 – 6 oblique or vertical septa, thick-walled, surface roughened, 18 – 20(– 21) × 7(– 8) µm. Culture characteristics — Colonies with abundant aerial mycelium and smooth, even margin, covering dish after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse iron-grey. Typus. south AfricA, Western Cape Province, Nieuwoudtville, on leaves of Leipoldtia schultzei (Aizoaceae), 2018, P.W. Crous, HPC 3024 (holotype CBS H-24418, culture ex-type CPC 38531 = CBS 146774, ITS, LSU and tub2 sequences GenBank MW175346.1, MW175386.1 and MW173137.1, MycoBank MB837838). Notes — Neocamarosporium was established for a genus of camarosporium-like fungi occurring on dying leaves of a Mesembryanthemum sp. (Aizoaceae) (Crous et al. 2014). Neocamarosporium leipoldtiae was collected in the same area, again occurring on a member of the Aizoaceae. Phylogenetically, however, it is closely related to Neocamarosporium salicorniicola, described from Salicornia sp. (Amaranthaceae) collected in Thailand (Wanasinghe et al. 2017). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Neocamarosporium sp. (strain CF-288928, GenBank MG065823.1; Identities = 544 /554 (98 %), one gap (0 %)), Pleosporales sp. 6 PV-2016 (strain DW, GenBank KU933734.1; Identities = 534/544 (98 %), two gaps (0 %)), and Neocamarosporium salicornicola (strain ZMCS3, GenBank MK809918.1; Identities = 509 /520 (98 %), two gaps (0 %)). Closest hits using the LSU sequence are Neocamarosporium chichastianum (strain CBS 137502, GenBank KP004483.1; Identities = 848 /853 (99 %), no gaps), Neocamarosporium salicornicola (strain MFLUCC 15-0957, GenBank MF434281.1; Identities = 842 /848 (99 %), no gaps), and Chaetosphaeronema hispidulum (strain CBS 826.88, GenBank EU754145.1; Identities = 862/870 (99 %), no gaps). Closest hits using the tub2 sequence had highest similarity to Neocamarosporium calvescens (strain T77I1, GenBank MK140511.1; Identities = 265/289 (92 %), one gap (0 %)), Phoma betae (strain CBS 109410, GenBank MK255063.1; Identities = 284 /311 (91 %), seven gaps (2 %)) and Dimorphosporicola tragani (strain CBS 570.85, GenBank KU728616.1; Identities = 412 /478 (86 %), 22 gaps (4 %)). Colour illustrations. Flowers of Leipoldtia schultzei. Conidiomata on MEA with central ostiole; conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 296 Persoonia – Volume 45, 2020 Dothiora aloidendri & Hantamomyces aloidendri 297 Fungal Planet description sheets Fungal Planet 1125 & 1126 – 19 December 2020 Dothiora aloidendri Crous, sp. nov. Etymology. Name refers to the host genus Aloidendron from which it was isolated. Classification — Dothioraceae, Dothideales, Dothideomycetes. Conidiomata pycnidial, globose, black, glabrous, erumpent, 200 – 350 µm diam, aggregated in dense clusters, forming a superficial layer on agar, exuding a creamy conidial mass. Conidiophores reduced to conidiogenous cells lining the inner cavity, hyaline, smooth, ampulliform to doliiform, phialidic, 6 – 9 × 5 –7 µm. Conidia solitary, straight, subcylindrical, aseptate, guttulate, hyaline, smooth, thin-walled, apex obtuse, tapering at base to truncate hilum, 1–1.5 µm diam, (10 –)12 –13(–14) × (3–) 4 µm. Culture characteristics — Colonies flat, spreading, with sparse to moderate aerial mycelium and smooth, lobate margin, reaching 40 mm diam after 2 wk at 25 °C. On MEA surface sepia, reverse isabelline; on PDA surface iron-grey, reverse olivaceous-grey; on OA surface olivaceous-grey. Typus. south AfricA, Western Cape Province, Namaqualand, on leaves of Aloidendron dichotomum (Asphodeloideae), 2018, P.W. Crous, HPC 3039 (holotype CBS H-24419, culture ex-type CPC 38535 = CBS 146775, ITS, LSU, tef1 and tub2 sequences GenBank MW175347.1, MW175387.1, MW173123.1 and MW173138.1, MycoBank MB837839). Notes — Species of Dothiora commonly form Dothichiza and hormonema-like morphs in culture (Crous & Groenewald 2016, 2017), as observed in D. aloidendri. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Dothiora europaea (strain EXF-12400, GenBank MK460357.1; Identities = 448/469 (96 %), two gaps (0 %)), Dothiora sorbi (strain CBS 742.71, GenBank KU728514.1; Identities = 554/591 (94 %), four gaps (0 %)), and Dothiora elliptica (strain CBS 736.71, GenBank KU728502.1; Identities = 489 /522 (94 %), two gaps (0 %)). Closest hits using the LSU sequence are Dothiora infuscans (strain FMR 16326, GenBank NG_066397.1; Identities = 782/792 (99 %), no gaps), Dothiora oleae (strain CBS 472.69, GenBank MH871116.1; Identities = 844 /856 (99 %), no gaps), and Neophaeocryptopus cytisi (strain MFLUCC 14-0970, GenBank NG_059643.1; Identities = 826 /838 (99 %), one gap (0 %)). Closest hits using the tef1 sequence had highest similarity to Dothiora oleae (strain CBS 235.57, GenBank KU728587.1; Identities = 195/200 (98 %), one gap (0 %)), Dothiora viburnicola (strain CBS 274.72, GenBank KU728591.1; Identities = 195/203 (96 %), no gaps), and Dothiora bupleuricola (strain CBS 112.75, GenBank KU728579.1; Identities = 194/201 (97 %), three gaps (0 %)). Closest hits using the tub2 sequence had highest similarity to Dothiora phillyreae (strain CBS 473.69, GenBank KU728629.1; Identities = 503/618 (81 %), 21 gaps (3 %)), Dothiora maculans (strain CBS 299.76, GenBank KU728621.1; Identities = 470/569 (83 %), 27 gaps (4 %)), and Dothiora oleae (strain CBS 152.71, GenBank KU728625.1; Identities = 495 /609 (81 %), 32 gaps (5 %)). Hantamomyces Crous, gen. nov. Etymology. Name refers to the Hantam district where it was collected, Nieuwoudtville, Northern Cape Province, South Africa. Classification — Ophiocordycipitaceae, Hypocreales, Sordariomycetes. Conidiophores arising from superficial hyphae, erect, solitary, cylindrical, pale brown, smooth, branched, septate. Conidiogenous cells integrated, pale brown, smooth, subcylindrical; conidiophores with terminal conidiogenous region with den- ticulate loci and with separating cell leaving minute collarette; conidiogenous cells giving rise to next cell in zigzag fashion (sympodial), appearing like a drawn out rachis. Conidia hyaline, smooth, fusoid, tapering towards both ends to truncate hilum with minute marginal frill and conidia occurring in long, unbranched chains, forming a mucoid droplet with age. Type species. Hantamomyces aloidendri Crous. MycoBank MB837840. Hantamomyces aloidendri Crous, sp. nov. Etymology. Name refers to the host genus Aloidendron from which it was isolated. Mycelium consisting of hyaline, smooth, branched, septate, 2.5– 3.5 µm diam hyphae. Conidiophores arising from superficial hyphae, erect, solitary, cylindrical, pale brown, smooth, branched, septate, up to 200 µm tall, 2.5–3 µm diam. Conidiogenous cells integrated, pale brown, smooth, subcylindrical; conidiophores with terminal conidiogenous region with 1–2 denticulate loci, 1 × 1 µm, with separating cell leaving minute collarette; one locus in basal region above septum, and second locus if present below apical septum, 10 – 30 × 2.5 – 3 µm, but conidiogenous cells giving rise to next cell in zigzag fashion (sympodial), appearing like a drawn out rachis. Conidia hyaline, smooth, guttulate to Colour illustrations. Aloidendron dichotomum growing along the mountain ridge. Left column D. aloidendri. Conidiomata on SNA; conidiogenous cells giving rise to conidia; conidia. Right column Hantamomyces aloidendri. Conidiophores giving rise to chains of conidia; conidia. Scale bars = 10 µm. granular, fusoid, not to slightly constricted at median septum, tapering towards both ends to truncate hilum with minute marginal frill, scar 1 µm diam, at times somewhat darkened, with conidia occurring in long, unbranched chains, forming a mucoid droplet with age; conidial chains with conidia attached to one another by minute separating cell, with collarette developing at each end, 1– 3-septate, (15 –)17–18(– 20) × (3.5–)4(– 5) µm. Culture characteristics — Colonies erumpent, spreading, with moderate aerial mycelium and smooth, even margin, reaching 20 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface dirty white, reverse buff. Typus. south AfricA, Western Cape Province, Nieuwoudtville, on leaves of Aloidendron dichotomum (Asphodelaceae), 2018, P.W. Crous, HPC 3020 (holotype CBS H-24432, culture ex-type CPC 38655 = CBS 146814, ITS and LSU sequences GenBank MW175348.1 and MW175388.1, MycoBank MB837841). (for Notes see Supplementary material page FP1125 & 1126; and for tree on Supplemetary material page FP1141) Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 298 Persoonia – Volume 45, 2020 Suttonomyces cephalophylli Fungal Planet description sheets 299 Fungal Planet 1127 – 19 December 2020 Suttonomyces cephalophylli Crous, sp. nov. Etymology. Name refers to the host genus Cephalophyllum from which it was isolated. Classification — Massarinaceae, Pleosporales, Dothideomycetes. Conidiomata solitary, immersed in host tissue, pycnidial, globose, brown, 150 – 200 µm diam; wall of 3 – 6 layers of brown textura angularis. Conidiophores reduced to conidiogenous cells, hyaline, smooth, phialidic, 4 – 6 × 3 – 5 µm. Conidia solitary, aseptate, medium brown, thick-walled, verruculose to spikey, ellipsoid with bluntly rounded ends, (12 –)14 –16(–18) × (7–)8 –10(–12) µm. Culture characteristics — Colonies flat, spreading, with sparse aerial mycelium and lobate, smooth margin, reaching 30 mm diam after 2 wk at 25 °C. On MEA surface dirty white in centre, cinnamon in outer region and in reverse; on PDA surface and reverse isabelline to cinnamon; on OA surface cinnamon. Typus. south AfricA, Western Cape Province, Clanwilliam, Rocklands camping, on leaves of Cephalophyllum pilansii (Aizoaceae), 2018, P.W. Crous, HPC 3055 (holotype CBS H-24420, culture ex-type CPC 38541 = CBS 146787, ITS and LSU sequences GenBank MW175349.1 and MW175389.1, MycoBank MB837842). Notes — The present collection clusters with species of Suttonomyces (Wijayawardene et al. 2015), but is distinct from known species in the genus in that it lacks muriformly-septate conidia and paraphyses. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Suttonomyces rosae (strain MFLU 18-0112, GenBank NR_157548.1; Identities = 447/469 (95 %), three gaps (0 %)), Stagonospora bicolor (strain LF2, GenBank KX510131.1; Identities = 347/367 (95 %), three gaps (0 %)), and Stagonospora pseudopaludosa (strain CBS 136424, GenBank NR_137840.1; Identities = 348 /370 (94 %), four gaps (1 %)). Closest hits using the LSU sequence are Suttonomyces rosae (strain MFLU 18-0112, GenBank NG_059882.1; Identities = 807/811 (99 %), no gaps), Helminthosporium velutinum (strain L136, GenBank KY984355.1; Identities = 838 /850 (99 %), no gaps), and Helminthosporium tiliae (strain L89, GenBank KY984346.1; Identities = 837/ 850 (98 %), no gaps). Colour illustrations. Flower and leaves of Cephalophyllum pilansii. Conidiomata on host tissue and on OA (scale bars = 200 µm); conidia (scale bar = 10 µm). Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 300 Persoonia – Volume 45, 2020 Endoconidioma euphorbiae 301 Fungal Planet description sheets Fungal Planet 1128 – 19 December 2020 Endoconidioma euphorbiae Crous, sp. nov. Etymology. Name refers to the host genus Euphorbia from which it was isolated. Classification — Dothioraceae, Dothideales, Dothideomycetes. Conidiomata erumpent, globose, black, pycnidial, 200–250 µm diam, with central ostiole exuding a black mucoid conidial mass. Conidiophores reduced to conidiogenous cells lining the inner cavity, hyaline, smooth, doliiform to ampulliform, 7–10 × 5 –7 µm, proliferating percurrently at apex. Conidia solitary, aseptate, golden-brown, thick-walled, verruculose, ellipsoid, apex obtuse, base bluntly rounded, (11–)12 –13(–14) × (7–)8(–9) µm. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and lobate, feathery margin, covering dish after 2 wk at 25 °C. On MEA and PDA surface and reverse iron-grey; on OA surface olivaceous grey. Typus. south AfricA, Western Cape Province, Nieuwoudtville, on leaves with tip dieback of Euphorbia mauritanica (Euphorbiaceae), 2018, P.W. Crous, HPC 3069 (holotype CBS H-24422, culture ex-type CPC 38551 = CBS 146776, ITS and LSU sequences GenBank MW175350.1 and MW175390.1, MycoBank MB837843). Additional material examined. south AfricA, Western Cape Province, Nieuwoudtville, on leaves with dieback of Brunsvigia bosmaniae (Amaryllidaceae), 2018, P.W. Crous, HPC 3041 (CBS H-24423, culture CPC 38583 = CBS 146777, ITS and LSU sequences GenBank MW175351.1 and MW175391.1). Endoconidioma Tsuneda et al., Mycologia 96: 1129. 2004. Synonym. Coniozyma Crous, In: Marincowitz et al., CBS Diversity Ser. (Utrecht) 7: 97. 2008. Endoconidioma carpetanum (Bills et al.) Crous, comb. nov. MycoBank MB837886 Basionym. Hormonema carpetanum Bills et al., Stud. Mycol. 50: 152. 2004. Endoconidioma leucospermi (Crous & Denman) Crous, comb. nov. MycoBank MB837887 Notes — Endoconidioma (based on E. populi ) is a genus originally described from twigs of Populus tremuloides collected in Canada. It is characterised by having a yeast-like morph in culture, as well as endoconidia, and a coelomycetous, coniothyrium-like morph (Tsuneda et al. 2004). Endoconidioma appears to be the oldest name for a clade in the Dothioraceae containing species with highly adaptable morphology. Coniozyma (based on C. leucospermi), is a morphologically highly variable fungus associated leaf spots of Proteaceae (Taylor & Crous 2001, Marincowitz et al. 2008), which appears to be better accommodated in Endoconidioma. Endoconidioma euphorbiae is phylogenetically related to E. leucospermi (conidia 6 –12 × 3 – 8 µm in vivo, 6.5 –15 × 3.5 – 8 µm in vitro; Taylor & Crous 2001), but distinguished based on its slightly larger conidia. Isolates from Brunsvigia bosmaniae (CPC 38583, conidia (13–)15–16(–17) × 7(–8) µm) are similar in size, though slightly more subcylindrical, and olivaceous brown in colour, but are accepted as falling within the variation for E. euphorbiae. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence of CPC 38551 had highest similarity to Coniozyma leucospermi (strain CBS 111289, GenBank EU552113.1; Identities = 579/592 (98 %), four gaps (0 %)), Hormonema carpetanum (strain 235J14, GenBank KU516485.1; Identities = 561/574 (98 %), four gaps (0 %)), and Endoconidioma populi (strain NWHC 46379-1433_1SD, GenBank MK782233.1; Identities = 516/528 (98 %), three gaps (0 %)). The ITS sequences of CPC 38551 and 38583 differ at two nucleotide positions (590/592 similar nucleotides). Closest hits using the LSU sequence of CPC 38551 are Coniozyma leucospermi (strain CBS 111289, GenBank EU552113.1; Identities = 844/849 (99 %), no gaps), Hormonema carpetanum (strain ATCC 74360, GenBank MF611880.1; Identities = 843/849 (99 %), no gaps), and Endoconidioma populi (strain UAMH 10297, GenBank NG_059198.1; Identities = 812/819 (99 %), no gaps). The LSU sequences of CPC 38551 and 38583 differ at one nucleotide position (811/812 similar nucleotides). Basionym. Coniothyrium leucospermi Crous & Denman, S. Afr. J. Bot. 64: 139. 1998. Synonym. Coniozyma leucospermi (Crous & Denman) Crous, In: Marincowitz et al., CBS Diversity Ser. (Utrecht) 7: 97. 2008. Colour illustrations. Euphorbia mauritanica. Conidia on SNA, and conidiomata oozing dark brown conidia on PNA (scale bars = 200 µm); conidiogenous cells giving rise to conidia; conidia (scale bars = 10 µm). (for tree see Supplemetary material page FP1141) Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 302 Persoonia – Volume 45, 2020 Neometulocladosporiella seifertii 303 Fungal Planet description sheets Fungal Planet 1129 – 19 December 2020 Neometulocladosporiella seifertii Crous, sp. nov. Etymology. Named after Keith A. Seifert, a Canadian mycologist who is always impressed by hyphomycetes with such magnificent, pigmented, solitary conidiophores. Classification — Rutstroemiaceae, Helotiales, Leotiomycetes. Conidiophores dimorphic. Microconidiophores erect, medium brown, smooth, solitary, subcylindrical, straight to flexuous, 1– 3-septate, 35 –70 × 3 – 5 µm, giving rise to a single, terminal conidiogenous cell. Conidiogenous cells 10 – 30 × 3 – 4 µm, medium brown, smooth, clavate, with a flat-tipped apical locus, 1– 2 µm diam, unthickened, not darkened, giving rise to ramoconidia. Macroconidiophores solitary, erect, straight to flexuous, unbranched, subcylindrical, dark brown, smooth, arising from superficial mycelium, 250–600 × 10–16 µm, 5–12-septate, dark brown, smooth, clavate, giving rise to a series of branches, 10–15 × 5 –7 µm, which are medium brown, smooth, subcylindrical to clavate, aseptate, base abruptly tapered to flat-tipped locus, 2 µm diam, apex with 2 – 4 denticles, 1 × 1 µm, unthickened, not darkened, giving rise to secondary ramoconidia. Primary ramoconidia fusoid-ellipsoid to subcylindrical or clavate, medium brown, smooth, 0 –1-septate, 10 – 22 × 5 – 6 µm, with 1– 3 apical flat-tipped loci, 1 µm diam, unthickened, not darkened. Secondary ramoconidia straight, pale brown, smooth to finely verruculose, 0 –1-septate, subcylindrical with obtuse ends, 10 –14 × 5 – 6 µm, base with abrupt taper to truncate hilum, 1–1.5 µm diam, apex with 1– 3 denticles, 1 µm diam, not thickened nor darkened, giving rise to branched, dry chains of acropetal conidia, pale brown, smooth to finely verruculose, subcylindrical to ellipsoid with obtuse ends, constricted at median septum, (8 –) 9 –10(–12) × (4 –) 4.5 (– 5) µm, with a flattipped basal hilum and 1– 3 apical denticles, 0.5 –1 µm diam, not thickened nor darkened. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and smooth, lobate margin, reaching 35 mm diam after 2 wk at 25 °C. On MEA surface isabelline, reverse hazel; on PDA surface isabelline, reverse umber; on OA surface isabelline. Notes — The hitherto monotypic genus Neometulocladosporiella was established for a genus of hyphomycetes occurring on Eucalyptus leaves collected in Colombia (Crous et al. 2018c). Morphologically the two species are very similar regarding their conidiophores, branches and conidial dimensions, and they are best distinguished based on the DNA sequence data. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Neometulocladosporiella eucalypti (strain CPC 31787, GenBank NR_160350.1; Identities = 535/555 (96 %), one gap (0 %)), Lanzia allantospora (strain CBS 124334, GenBank AB926099.1; Identities = 527/559 (94 %), nine gaps (1 %)), and Rutstroemia firma (voucher TU 104487, GenBank LT158448.1; Identities = 516/555 (93 %), nine gaps (1 %)). Closest hits using the LSU sequence are Neometulocladosporiella eucalypti (strain CPC 31787, GenBank NG_064541.1; Identities = 831/836 (99 %), no gaps), Lanzia allantospora (strain CBS 124334, GenBank AB926154.1; Identities = 838/847 (99 %), no gaps), and Ciboria americana (voucher KUS-F52240, GenBank JN086702.1; Identities = 744/760 (98 %), no gaps). Typus. south AfricA, Western Cape Province, Clanwilliam, on leaves of Combretum caffrum (Combretaceae), 2018, P.W. Crous, HPC 3048 (holotype CBS H-24424, culture ex-type CPC 38599 = CBS 146795, ITS and LSU sequences GenBank MW175352.1 and MW175392.1, MycoBank MB837844). Colour illustrations. Leaves and branches of Combretum caffrum. Erect mononematous macroconidiophores on SNA; Clavate conidiophores giving rise to a series of branches and conidiogenous cells; microconidiophores and chains of conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 304 Persoonia – Volume 45, 2020 Verrucocladosporium carpobroti 305 Fungal Planet description sheets Fungal Planet 1130 – 19 December 2020 Verrucocladosporium carpobroti Crous, sp. nov. Etymology. Name refers to the host genus Carpobrotus from which it was isolated. Classification — Cladosporiaceae, Cladosporiales, Dothideomycetes. Conidiophores solitary, erect, straight to flexuous, branched, subcylindrical, medium brown, verruculose, arising from superficial mycelium, 50 – 200 × 5 – 6 µm, 2 –10-septate, giving rise to a series of branches, 20 – 50 × 5 – 6 µm, which are medium brown, verruculose, subcylindrical, 1– 3-septate. Conidiogenous cells integrated, subcylindrical, medium brown, verruculose, terminal and intercalary, 20 – 40 × 3 – 5 µm; loci thickened, darkened and refractive, 2 – 3 µm diam. Primary ramoconidia fusoid-ellipsoid to subcylindrical, thick-walled, medium brown, verruculose to warty, 0 – 2-septate, 25 – 55 × 4–5 µm, with 1–3 apical, flat-tipped loci, 2 µm diam, thickened, darkened. Secondary ramoconidia straight, medium brown, verruculose to warty, thick-walled, 0 –1-septate, subcylindrical to fusoid-ellipsoid, 15–20 × 4–5 µm; hila thickened and darkened, 1.5–2 µm diam, giving rise to branched, dry chains of acropetal conidia, medium brown, verruculose to warty, subcylindrical to fusoid-ellipsoid, 0(–1)-septate, (10 –)12 –14(–16) × (4 –)5 – 6 µm; hila 1.5 – 2 µm diam, thickened and darkened. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and smooth, even margin, reaching 20 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse olivaceous grey. Notes — Verrucocladosporium was introduced to accommodate cladosporium-like species having ± planate, non-coronate conidiogenous loci and hila, and warty, verrucose conidia. Verrucocladosporium carpobroti is related to V. dirinae (conidiophores up to 85 µm long, conidia 4 –18 (–23) × (2 –)2.5 – 3.5 µm, 0 –1-septate; Crous et al. 2007a), but is morphologically distinct. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Verrucocladosporium dirinae (strain CBS 112794, GenBank NR_152317.1; Identities = 496/509 (97 %), no gaps), Verrucocladosporium visseri (strain CPC 36317, GenBank NR_166320.1; Identities = 475 /487 (98 %), one gap (0 %)), and Graphiopsis chlorocephala (strain CPC 11969, GenBank EU009458.2; Identities = 475 /498 (95 %), six gaps (1 %)). Closest hits using the LSU sequence are Verrucocladosporium dirinae (strain MUT<ITA> 4857, GenBank KP671739.1; Identities = 865/873 (99 %), no gaps), Graphiopsis chlorocephala (strain CPC 11969, GenBank EU009458.2; Identities = 865/873 (99 %), no gaps), and Verrucocladosporium visseri (strain CPC 36317, GenBank NG_068322.1; Identities = 861/869 (99 %), no gaps) Typus. south AfricA, Western Cape Province, Clanwilliam, on leaves of Carpobrotus quadrifidus (Aizoaceae), 2018, P.W. Crous, HPC 3027 (holotype CBS H-24427, culture ex-type CPC 38635 = CBS 146784, ITS and LSU sequences GenBank MW175353.1 and MW175393.1, MycoBank MB837845). Additional material examined. south AfricA, Western Cape Province, Namaqualand, on leaves of Dimorphotheca sp. (Asteraceae), 2018, P.W. Crous, HPC 3040 (CBS H-24430, culture CPC 38645 = CBS 146796, ITS and LSU sequences GenBank MW175354.1 and MW175394.1). Colour illustrations. Flower of Carpobrotus quadrifidus. Conidiophores on SNA; conidiogenous cells giving rise to conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 306 Persoonia – Volume 45, 2020 Stemphylium carpobroti 307 Fungal Planet description sheets Fungal Planet 1131 – 19 December 2020 Stemphylium carpobroti Crous, sp. nov. Etymology. Name refers to the host genus Carpobrotus from which it was isolated. Classification — Pleosporaceae, Pleosporales, Dothideomycetes. Mycelium consisting of brown, septate, branched, finely verruculose, 3 – 4 µm diam hyphae. Conidiophores solitary, erect, subcylindrical, mostly unbranched, brown, finely verruculose, 40–120 × 4–7 µm, 3–5-septate, becoming swollen towards conidiogenous cell. Conidiogenous cells terminal, clavate, brown, finely verruculose, thick-walled, 10–20 × 8–9 µm, with terminal locus, 3–4 µm diam. Conidia solitary, dark brown, verruculose, ellipsoid to obovoid, constricted at medium septum, tapering to subobtuse apex, (30 –) 35 – 45(–70) × (17–) 20 – 25 µm, with (3 –) 4 (– 6) transverse septa, and 1– 4 vertical or oblique septa per transverse section. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and smooth, even margin, reaching 50 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse iron-grey. Typus. south AfricA, Western Cape Province, Clanwilliam, on leaves of Carpobrotus quadrifolius (Aizoaceae), 2018, P.W. Crous, HPC 3027 (holotype CBS H-24428, culture ex-type CPC 38637 = CBS 146789, ITS, LSU and gapdh sequences GenBank MW175355.1, MW175395.1 and MW173103.1, MycoBank MB837846). Notes — Stemphylium carpobroti is closely related to S. novae-zelandiae (conidia (31–)34 – 40.5(– 45.5) × (9 –)11– 13(–14.5) μm, with 3 – 5(–7) transverse septa and 1– 2 longitudinal or oblique septa per transverse sector; Woudenberg et al. 2017), but is distinct in having larger conidia. Stemphylium vesicarium is also closely related, but generally has shorter conidia (see Woudenberg et al. 2017). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Asteromyces cruciatus (strain CBS 171.63, GenBank NR_159604.1; Identities = 548 /564 (97 %), six gaps (1 %)), Stemphylium vesicarium (strain NIHHS404, GenBank KY555005.1; Identities = 553 /571 (97 %), four gaps (0 %)), and Stemphylium lucomagnoense (strain CIRM-BRFM2667, GenBank MK691703.1; Identities = 560/579 (97 %), four gaps (0 %)). Closest hits using the LSU sequence are Stemphylium botryosum (strain CBS 714.68, GenBank NG_069738.1; Identities = 849 /851 (99 %), no gaps), Stemphylium vesicarium (strain 18ALIM004, GenBank MT472605.1; Identities = 849/851 (99 %), no gaps), and Stemphylium eturmiunum (strain CBS 109845, GenBank NG_069866.1; Identities = 842/844 (99 %), no gaps). Closest hits using the gapdh sequence had highest similarity to Stemphylium lycopersici (strain G9RS, GenBank MN393479.1; Identities = 369/377 (98 %), no gaps), Stemphylium vesicarium (strain On16-499, GenBank MK675745.1; Identities = 369 /377 (98 %), no gaps), and Stemphylium globuliferum (strain SWp202, GenBank KF479194.1; Identities = 369 /377 (98 %), no gaps). Colour illustrations. Leaves of Carpobrotus quadrifolius. Conidiophores and conidiogenous cells giving rise to conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 308 Persoonia – Volume 45, 2020 Neocladosporium osteospermi 309 Fungal Planet description sheets Fungal Planet 1132 – 19 December 2020 Neocladosporium osteospermi Crous, sp. nov. Etymology. Name refers to the host genus Osteospermum from which it was isolated. Classification — Cladosporiaceae, Cladosporiales, Dothideomycetes. Mycelium of branched, septate, 2.5 – 3 µm diam hyphae, mot constricted at septa, medium brown, verruculose. Conidiophores reduced to conidiogenous cells on hyphae, or erect, straight, sometimes slightly flexuous, narrowly cylindrical, non-geniculate, or nodulose, unbranched, 0 – 2-septate, up to 65 µm long, 2–3 µm wide, medium brown, verruculose. Conidiogenous cells integrated, mostly terminal, sometimes intercalary, cylindrical, 15 – 35 µm long, proliferating sympodially with 1– 3 conidiogenous loci, 2–3 µm diam, thickened, darkened and refractive. Ramoconidia cylindrical, 15–35 × 4–5 µm, 1–3-septate, concolorous with conidiophores, thick-walled, irregularly rough-walled, smooth to verruculose to warty, apically with up to two hila, 2 – 3 µm diam, thickened, darkened and refractive. Conidia catenate, in branched, chains, ellipsoid, fusoid to subcylindrical, (11–)13 –16(– 20) × (3 –) 3.5 (– 4) µm, 0 –1-septate, medium brown, thick-walled, smooth to verruculose to warty, somewhat attenuated towards both ends, hila truncate, 2–3 µm diam, darkened, thickened and refractive. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and smooth, even margin, reaching 35 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse olivaceous grey. Notes — Neocladosporium presently contains two species, namely N. leucadendri and N. syringae, characterised by having conidia with a warty, mucoid outer layer (Bezerra et al. 2017, Crous et al. 2020b). Neocladosporium osteospermi adds a third species to the genus. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Neocladosporium syringae (strain CPC 35750, GenBank NR_170057.1; Identities = 648/681 (95 %), 13 gaps (1 %)), Davidiellomyces australiensis (strain CBS 142165, GenBank NR_154036.1; Identities = 612/687 (89 %), 22 gaps (3 %)), and Davidiellomyces juncicola (strain CPC 38038, GenBank NR_166347.1; Identities = 618/699 (88 %), 28 gaps (4 %)). Closest hits using the LSU sequence are Neocladosporium syringae (strain CPC 35750, GenBank MT223912.1; Identities = 774/778 (99 %), one gap (0 %)), Neocladosporium leucadendri (strain CBS 131317, GenBank NG_057949.1; Identities = 833/841 (99 %), no gaps), and Neocladosporium leucadendri (as Toxicocladosporium leucadendri; strain CPC 29092, GenBank LT799745.1; Identities = 738/746 (99 %), no gaps). Typus. south AfricA, Western Cape Province, Clanwilliam, on leaf spots of Osteospermum moniliferum (Asteraceae), 2018, P.W. Crous, HPC 3035 (holotype CBS H-24429, culture ex-type CPC 38641 = CBS 146813, ITS and LSU sequences GenBank MW175356.1 and MW175396.1, MycoBank MB837847). Colour illustrations. Flower of Osteospermum moniliferum. Conidiophores and conidiogenous cells giving rise to conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 310 Persoonia – Volume 45, 2020 Eucasphaeria proteae 311 Fungal Planet description sheets Fungal Planet 1133 – 19 December 2020 Eucasphaeria proteae Crous, sp. nov. Etymology. Name refers to the host genus Protea from which it was isolated. Classification — Niessliaceae, Hypocreales, Sordariomycetes. Mycelium consisting of hyaline, smooth, branched, septate, 1.5–2.5 µm diam hyphae. Conidiomata sporodochial, 100–300 µm diam, becoming aggregated, forming large orange, mucoid colonies on agar; basal stroma of hyaline textura angularis, giving rise to hyaline, smooth, branched, 3 –10-septate conidiophores, subcylindrical, 20–70 × 2–3 µm. Conidiogenous cells integrated, terminal and intercalary, subcylindrical, flexuous, phialidic, hyaline, smooth, 8 – 20 × 2.5 – 3 µm. Conidia solitary, hyaline, smooth, guttulate, subcylindrical, straight to slightly curved, apex obtuse, base truncate, 1.5–2 µm diam, 0(–3)-septate, (8 –)15 –17(– 20) × (2 –)2.5(– 3) µm; 3-septate conidia can become up to 65 µm in length, and frequently undergo microcyclic conidiation. Culture characteristics — Colonies erumpent, spreading, with sparse aerial mycelium and smooth, lobate margin, reaching 25 mm diam after 2 wk at 25 °C. On MEA, PDA and OA, surface and reverse orange. Typus. south AfricA, Western Cape Province, Clanwilliam, on leaves of Protea neriifolia (Proteaceae), 2018, P.W. Crous, HPC 3030 (holotype CBS H-24433, culture ex-type CPC 38661 = CBS 146815, ITS, LSU, rpb2 and tef1 (second part) sequences GenBank MW175357.1, MW175397.1, MW173116.1 and MW173129.1, MycoBank MB837848). Notes — Culture CPC 38661 was originally derived from hyaline, aseptate microconidia found on the surface of leaves of Protea neriifolia. In culture, sporodochia with an Eucasphaeria asexual morph developed (Crous et al. 2007b). Based on LSU the present fungus proved to be closely related to Rosasphaeria moravica, which was described as forming densely aggregated orange pycnidial conidiomata in culture (Jaklitsch & Voglmayr 2012). Based on the sporodochia, and conidial morphology, the present collection is best accommodated in Eucasphaeria, although the relationship with Rosasphaeria deserves further study. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Rosasphaeria moravica (strain CBS 124270, GenBank NR_138377.1; Identities = 489/572 (85 %), 26 gaps (4 %)), Neoeucasphaeria eucalypti (strain CBS 145075, GenBank NR_161136.1; Identities = 489/573 (85 %), 35 gaps (6 %)), and Eucasphaeria rustici (strain CPC 28946, GenBank NR_154028.1; Identities = 488/573 (85 %), 31 gaps (5 %)). Closest hits using the LSU sequence are Rosasphaeria moravica (strain LMM, GenBank JF440985.1; Identities = 836/851 (98 %), two gaps (0 %)), Eucasphaeria capensis (strain CBS 120027, GenBank EF110619.1; Identities = 855/874 (98 %), two gaps (0 %)), and Niesslia pulchriseta (strain CBS 839.96, GenBank MG826846.1; Identities = 854/877 (97 %), five gaps (0 %)). Closest hits using the rpb2 sequence had highest similarity to Rosasphaeria moravica (strain LMM, GenBank JF440986.1; Identities = 586/683 (86 %), two gaps (0 %)), Ophiocordyceps mosingtoensis (strain BCC 30904, GenBank MK214100.1; Identities = 545/679 (80 %), six gaps (0 %)), and Ophiocordyceps coccidiicola (strain NBRC 100682, GenBank AB968545.1; Identities = 542/678 (80 %), two gaps (0 %)). Closest hits using the tef1 (second part) sequence had highest similarity to Tolypocladium tropicale (strain MX338, GenBank KF747113.1; Identities = 793 /864 (92 %), two gaps (0 %)), Isaria takamizusanensis (strain F896, GenBank GU979994.1; Identities = 831/911 (91 %), two gaps (0 %)), and Nectria haematococca (strain GJS89-70, GenBank AY489624.1; Identities = 831/912 (91 %), four gaps (0 %)). Colour illustrations. Flowers and leaves of Protea neriifolia. Sporodochia on OA; conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 312 Persoonia – Volume 45, 2020 Paramyrothecium pituitipietianum 313 Fungal Planet description sheets Fungal Planet 1134 – 19 December 2020 Paramyrothecium pituitipietianum Crous, sp. nov. Etymology. Composed of pituita (= mucus, snot) and the name Piet (referring to ‘Pietsnot’ = Snotty Pete, the common South African name of Grielum humifusum). Classification — Stachybotryaceae, Hypocreales, Sordariomycetes. Conidiomata sporodochial, stromatic, superficial, cupulate, separate to gregarious, oval, 200 – 350 µm diam with a white, setose fringe surrounding the dark green mucoid conidial mass. Stroma well-developed of hyaline textura angularis. Setae thickwalled, 7–10-septate, straight to flexuous, hyaline, 100 – 300 × 4 – 5 µm, tapering to obtuse apex. Conidiophores penicillately branched, hyaline, smooth, 2 – 4-septate, 20 – 35 × 3 – 4 µm. Conidiogenous cells phialidic, hyaline, smooth, subcylindrical, tapering at tip, 10 –15 × 2 – 2.5 µm. Conidia aseptate, subcylindrical, straight, pale green, smooth, guttulate, ends obtuse, (7–)9 –10(–12) × (2–)2.5 µm. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and smooth, even margin, reaching 35 mm diam after 2 wk at 25 °C. On MEA surface folded, buff, reverse luteous; on PDA surface and reverse buff; on OA surface cinnamon. Typus. south AfricA, Western Cape Province, Nieuwoudtville, on stems of Grielum humifusum (Neuradaceae), 2018, P.W. Crous, HPC 3057 (holotype CBS H-24435, culture ex-type CPC 38688 = CBS 146817, ITS, LSU, cmdA, tef1 and tub2 sequences GenBank MW175358.1, MW175398.1, MW173100.1, MW173124.1 and MW173139.1, MycoBank MB837849). Notes — Lombard et al. (2016) distinguished Paramyrothecium from Myrothecium s.str. and the other myrothecium-like genera by their septate, thin-walled setae surrounding the sporodochia. Paramyrothecium pituitipietianum is closely related to P. parvum (conidia 4 – 5 × 1– 2 µm) and P. telicola (conidia (7–)7.5 – 8.5(– 9) × 1– 3 μm; Lombard et al. 2016). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Paramyrothecium parvum (strain CBS 257.35, GenBank NR_145076.1; Identities = 577/589 (98 %), one gap (0 %)), Paramyrothecium roridum (as Myrothecium roridum; strain BBA 62764, GenBank AJ301993.1; Identities = 589/602 (98 %), one gap (0 %)), and Paramyrothecium acadiense (strain CBS 123.96, GenBank KU846288.1; Identities = 576/589 (98 %), one gap (0 %)). Closest hits using the LSU sequence are Paramyrothecium nigrum (strain CBS 116537, GenBank NG_069341.1; Identities = 826/827 (99 %), no gaps), Paramyrothecium foliicola (strain CBS 419.93, GenBank KU846323.1; Identities = 826/827 (99 %), no gaps), and Paramyrothecium roridum (strain CBS 372.50, GenBank MH868182.1; Identities = 845/847 (99 %), no gaps). Closest hits using the cmdA sequence had highest similarity to Paramyrothecium tellicola (strain CBS 478.91, GenBank KU846272.1; Identities = 488/600 (81 %), 17 gaps (2 %)), Paramyrothecium sinense (strain ZSY8, GenBank MH885437.1; Identities = 470/578 (81 %), 19 gaps (3 %)), and Xepicula crassiseta (strain CBS 392.71, GenBank KU847222.1; Identities = 494/608 (81 %), 30 gaps (4 %)). Distant hits using the tef1 sequence had highest similarity to Neomyrothecium humicola (strain CBS 310.96, GenBank KU846527.1; Identities = 207/229 (90 %), six gaps (2 %)), Gregatothecium humicola (strain CBS 205.96, GenBank KU846402.1; Identities = 212/237 (89 %), ten gaps (4 %)), and Brevistachys ossiformis (strain CPC 16031, GenBank KU846092.1; Identities = 209/234 (89 %), nine gaps (3 %)). Closest hits using the tub2 sequence had highest similarity to Paramyrothecium sp. 2 MP-2020 (strain 18ALOM016, GenBank MT671910.1; Identities = 324/330 (98 %), no gaps), Paramyrothecium terrestris (strain CBS 564.86, GenBank KU846420.1; Identities = 310/343 (90 %), seven gaps (2 %)), and Paramyrothecium acadiense (strain CBS 123.96, GenBank KU846405.1; Identities = 308/342 (90 %), six gaps (1 %)). Colour illustrations. Characteristic yellow flowers of Grielum humifusum. Conidioma on PNA (scale bar = 350 µm); setae; conidiogenous cells; conidia (scale bars = 10 µm). Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 314 Persoonia – Volume 45, 2020 Polyscytalum pini-canariensis 315 Fungal Planet description sheets Fungal Planet 1135 – 19 December 2020 Polyscytalum pini-canariensis Crous, sp. nov. Etymology. Name refers to the host genus Pinus from which it was isolated. Classification — Phlogicylindriaceae, Xylariales, Sordariomycetes. Mycelium consisting of brown, smooth, septate, 2 – 3 µm diam hyphae. Conidiophores erect, solitary, subcylindrical, branched or not, brown, smooth, flexuous, 1–5-septate, 20–40 × 2–3 µm. Conidiogenous cells integrated, terminal and intercalary, 10–20 × 2 – 3 µm, proliferating sympodially, denticulate, flat-tipped, 2.5–3 µm diam, not thickened nor darkened. Conidia occurring in unbranched chains, cylindrical with truncate ends, smooth, guttulate, medianly 1-septate, (18–)22–26(–48) × 3(–3.5) µm. Culture characteristics — Colonies erumpent, spreading, with moderate aerial mycelium and smooth, even margin, reaching 16 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse isabelline. Typus. spAiN, Canary Islands, Gran Canaria, N28°3'18" O15°41'43", 520 m, on needles of Pinus canariensis (Pinaceae), 4 July 2019, J. Etayo, HPC 3084 (holotype CBS H-24437, culture ex-type CPC 38727 = CBS 146819, ITS, LSU and actA sequences GenBank MW175359.1, MW175399.1 and MW173095.1, MycoBank MB837850). Notes — Polyscytalum pini-canariensis should be compared to P. pini (on Pinus sylvestris, UK; conidia (0 –)1(– 2)-septate, 7–12(–14) × 1.5 – 2(– 2.5) μm, conidiophores 50 –110(–140) μm; Kirk 1983 and P. pinicola (on Pinus tecunumanii, Malaysia; conidia (0–)1-septate, (13–)14–15(–16) × 2 μm, conidiophores 40–80 × 2–3 μm; Crous et al. 2020b). The new species can be distinguished based on its shorter conidiophores, and longer conidia. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Polyscytalum neofecundissimum (strain CBS 143390, GenBank NR_158959.1; Identities = 548/590 (93 %), 12 gaps (2 %)), Subulispora britannica (strain ICMP 14767, GenBank EF029198.1; Identities = 533/585 (91 %), 16 gaps (2 %)), and Polyscytalum pinicola (strain CPC 36759, GenBank MT223833.1; Identities = 539 /606 (89 %), 15 gaps (2 %)). Closest hits using the LSU sequence are Polyscytalum fecundissimum (strain CBS 100506, GenBank EU035441.1; Identities = 792/801 (99 %), one gap (0 %)), Polyscytalum chilense (strain CBS 143387, GenBank MH107954.1; Identities = 824/834 (99 %), one gap (0 %)), and Polyscytalum eucalyptigenum (strain CBS 143388, GenBank MH107955.1; Identities = 822/833 (99 %), one gap (0 %)). No significant hits were obtained when the actA sequence was used in blastn and megablast searches. Colour illustrations. Pinus canariensis covered in lichens growing on the Canary Islands. Conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl Javier Etayo, Department of Biology, IES Zizur, Ronda S. Cristóbol 196,31180 Zizur Mayor, Navarra, Spain; e-mail: jetayosa@educacion.navarra.es © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 316 Persoonia – Volume 45, 2020 Acremonium behniae 317 Fungal Planet description sheets Fungal Planet 1136 – 19 December 2020 Acremonium behniae Crous, sp. nov. Etymology. Name refers to the host genus Behnia from which it was isolated. Classification — Bionectriaceae, Hypocreales, Sordariomycetes. Mycelium consisting of hyaline, smooth, septate, branched, 1.5 – 2 µm diam hyphae. Conidiophores reduced to conidiogenous cells, erect, straight to flexuous, hyaline, smooth, phialidic, arising from superficial hyphae or from hyphal strands, giving rise to mucoid balls of conidia, but conidiogenous cells aggregated on hyphal strands, forming a sporodochial mass on agar surface conidiogenous cells subcylindrical with apical taper, 10 – 30 × 1.5– 2 µm; apex 1–1.5 µm diam, with minute non-flares collarette, 1 µm tall. Conidia hyaline, smooth, aseptate, subcylindrical to fusoid-ellipsoid, apex subobtuse, base bluntly rounded, (3.5 –)4 – 5(– 6.5) × 1.5– 2 µm. Culture characteristics — Colonies flat, spreading, with folded surface, moderate aerial mycelium and smooth, lobate margin, reaching 30 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface dirty white, reverse buff to dirty white. Notes — Acremonium behniae is closely related to A. charticola (conidiogenous cells 15 – 45(– 60) × 1.5 – 2(– 2.5) µm, conidia 3.2 – 4.5 × 1.4 – 2 µm; Gams 1971), but can be distinguished based on dimensions of its conidiogenous cells and conidia. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Acremonium charticola (strain UOA /HCPF 14413, GenBank KC253940.1; Identities = 525 /570 (92 %), nine gaps (1 %)) and Acremonium sclerotigenum (strain CBS 286.70H, GenBank MH859618.1; Identities = 535/581 (92 %), 14 gaps (2 %)). Closest hits using the LSU sequence are Acremonium sclerotigenum (strain UBOCC-A-118074, GenBank MT226553.1; Identities = 786 /789 (99 %), one gap (0 %)), Acremonium sordidulum (strain CBS 385.73, GenBank MH872418.1; Identities = 837/841 (99 %), no gaps), and Acremonium alternatum (strain CBS 407.66, GenBank FJ176883.1; Identities = 837/841 (99 %), no gaps). Typus. south AfricA, Northern Province, Tzaneen, Buffelskloof Nature Reserve, on leaves of Behnia reticulata (Asparagaceae), 2018, P.W. Crous, HPC 3156 (holotype CBS H- 24443, culture ex-type CPC 38798 = CBS 146824, ITS and LSU sequences GenBank MW175360.1 and MW175400.1, MycoBank MB837851). Colour illustrations. Leaves of Behnia reticulata. Sporulating colony on SNA; conidiophores and conidiogenous cells giving rise to conidia; conidia. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 318 Persoonia – Volume 45, 2020 Alternaria mirabibensis Fungal Planet description sheets 319 Fungal Planet 1137 – 19 December 2020 Alternaria mirabibensis Crous, sp. nov. Etymology. Name refers to the collection site, namely the Mirabib Rock in the Namib Desert, Namibia, where Stanley Kubrick filmed ‘the dawn of mankind’ in the movie ‘2001- A Space Odyssey’. Notes — Alternaria mirabibensis is closely related to Alternaria burnsii (CBS 130264) (Woudenberg et al. 2015, Nishikawa & Nakashima 2020), but is phylogenetically distinct. Classification — Pleosporaceae, Pleosporales, Dothideomycetes. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Alternaria alternata (strain KU20017.1, GenBank MT487794.1; Identities = 552/566 (98 %), two gaps (0 %)), Alternaria arborescens (strain ALT-14, GenBank MH879771.1; Identities = 552/566 (98 %), two gaps (0 %)), and Alternaria burnsii (strain CBS 130264, GenBank MH865506.1; Identities = 552 /566 (98 %), two gaps (0 %)). Closest hits using the LSU sequence are Alternaria multiformis (strain CBS 102060, GenBank NG_069860.1; Identities = 870/871 (99 %), no gaps), Alternaria terricola (strain CBS 202.67, GenBank NG_069728.1; Identities = 870/871 (99 %), no gaps), and Alternaria atra (strain CBS 125894, GenBank MH875550.1; Identities = 870/871 (99 %), no gaps). Closest hits using the actA sequence had highest similarity to Alternaria iridicola (strain AC139, GenBank LC481866.1; Identities = 179/185 (97 %), no gaps), Alternaria alternata (strain LSA2, GenBank KY131956.1; Identities = 186/193 (96 %), no gaps), and Alternaria tenuissima (strain U-2, GenBank MN752246.1; Identities = 208/216 (96 %), no gaps). Closest hits using the chs-1 sequence had highest similarity to Alternaria novae-guineensis (strain SCSJ08, GenBank MH793684.1; Identities = 233/242 (96 %), no gaps), Alternaria solani (strain NL03003, GenBank CP022032.1; Identities = 257/269 (96 %), no gaps), and Alternaria radicina (strain BMP0079, GenBank EU141977.1; Identities = 252/264 (95 %), no gaps). Closest hits using the cmdA sequence had highest similarity to Alternaria alstroemeriae (strain CBS 118809, GenBank MH175185.1; Identities = 579/639 (91 %), 15 gaps (2 %)), Alternaria iridiaustralis (strain CBS 118486, GenBank MH175191.1; Identities = 578 / 638 (91 %), 15 gaps (2 %)), and Alternaria alternata (strain 17MC, GenBank MG925134.1; Identities = 580 /647 (90 %), 25 gaps (3 %)). Closest hits using the gapdh sequence had highest similarity to Alternaria tenuissima (strain GP4, GenBank MK451969.1; Identities = 553 /577 (96 %), no gaps), Alternaria longipes (strain AXLKY2019010, GenBank MN044655.1; Identities = 552/577 (96 %), no gaps), and Alternaria alternata (strain D11, GenBank MK732570.1; Identities = 552/577 (96 %), no gaps). Closest hits using the tef1 sequence had highest similarity to Alternaria alternata (strain EGS 34-016, GenBank AH013339.2; Identities = 284 /304 (93 %), no gaps), Alternaria jacinthicola (strain Mlb684, GenBank HQ413697.1; Identities = 293/317 (92 %), no gaps), and Alternaria longipes (strain KY_2019_012, GenBank MT548042.1; Identities = 247/273 (90 %), five gaps (1 %)). Closest hits using the tub2 sequence had highest similarity to Alternaria arborescens (strain BAS_G1, GenBank MF070272.1; Identities = 282/289 (98 %), no gaps), Alternaria tenuissima (strain CBS 124278, GenBank MF070256.1; Identities = 282 /289 (98 %), no gaps), and Alternaria gaisen (strain CBS 118488, GenBank MF070254.1; Identities = 282 /289 (98 %), no gaps). Mycelium consisting of pale brown, smooth, branched, septate, 3 – 4 µm diam hyphae. Conidiophores erect, solitary, arising from superficial mycelium, 50 –150 × 3 – 5 µm, 3 – 6-septate, branched or not, brown, smooth, subcylindrical, straight to flexuous. Conidiogenous cells terminal and intercalary, straight to geniculous-sinuous, flexuous, 10–30 × 5–7 µm, with thickened, darkened, 1– 2 terminal pores, 2 – 3 µm diam. Conidia occurring in branched chains, conidia brown, verruculose, guttulate, ovoid to ellipsoid, (23 –)33 – 45(– 50) × (13 –)15 –16(–17) µm (body excluding beak), with 3 – 6 (–7) transverse septa, and (1–)2 – 3(– 6) longitudinal or oblique septa, commonly forming a long terminal beak, 20 –120 µm long, that becomes a secondary conidiophore, giving rise to terminal and lateral chains of conidia. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and smooth, lobate margin, reaching 40 mm diam on MEA, but covering dish on PDA and OA after 2 wk at 25 °C. On MEA surface folded, grey olivaceous, reverse isabelline; on PDA surface and reverse grey olivaceous; on OA surface grey olivaceous. Typus. NAmibiA, Gobabeb-Namib Research Institute, Mirabib, on plant litter, 19 Nov. 2019, P.W. Crous, HPC 3108 (holotype CBS H-24445, culture ex-type CPC 38838 = CBS 146826, ITS, LSU, actA, chs-1, cmdA, gapdh, tef1 and tub2 sequences GenBank MW175361.1, MW175401.1, MW173096.1, MW173101.1, MW173102.1, MW173104.1, MW173125.1 and MW173140.1, MycoBank MB837852). Colour illustrations. View from top of Mirabib Rock, looking outwards across the Namib Desert. Conidiophores and conidiogenous cells giving rise to conidial chains. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl Neriman Yilmaz, Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), Faculty of Natural and Agricultural Sciences, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: neriman.yilmazvisagie@fabi.up.ac.za Don A. Cowan, Centre for Microbial Ecology and Genomics, Department of Biochemistry, Genetics and Microbiology, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: don.cowan@up.ac.za Gillian Maggs-Kölling, Gobabeb-Namib Research Institute, P.O. Box 953, Walvis Bay, Namibia; e-mail: gillian@gobabeb.org © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 320 Persoonia – Volume 45, 2020 Preussia procaviae Fungal Planet description sheets 321 Fungal Planet 1138 – 19 December 2020 Preussia procaviae Crous, sp. nov. Etymology. Name refers to Procavia capensis (rock rabbit), from who’s dung this fungus was isolated. Classification — Sporormiaceae, Pleosporales, Dothideomycetes. Conidiomata pycnidial, solitary, immersed to erumpent, brown, globose, 60 –150 µm diam, with central ostiole, 10 µm diam; wall of 3 – 6 layers of brown textura angularis. Conidiophores reduced to conidiogenous cells lining the inner cavity, hyaline, smooth, phialidic, 4 – 5 × 2.5 – 3 µm. Conidia solitary, aseptate, hyaline, smooth, guttulate, ellipsoid with obtuse ends, 3 – 4 × 2 µm. Culture characteristics — Colonies flat, spreading, surface folded, with sparse aerial mycelium and smooth, lobate margin, reaching 50 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface olivaceous grey, reverse iron-grey. Typus. NAmibiA, Gobabeb-Namib Research Institute, Mirabib Rock, on dung of Procavia capensis (Procaviidae), 19 Nov. 2019, P.W. Crous, HPC 3110 (holotype CBS H-24446, culture ex-type CPC 38861 = CBS 146827, ITS, LSU, tef1 and tub2 sequences GenBank MW175362.1, MW175402.1, MW173126.1 and MW173141.1, MycoBank MB837853). Notes — The sexual morph of Preussia procaviae was not observed on the dung, nor did it develop in culture. However, species of Preussia are known to form phoma-like asexual morphs in culture. Preussia procaviae is phylogenetically distinct from its closest relatives. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Preussia sp. (strain CF209171, GenBank KX710223.1; Identities = 510/511 (99 %), one gap (0 %)), Sporormiella intermedia (as Preussia intermedia; strain OK2L126P, GenBank KF871451.1; Identities = 476/494 (96 %), four gaps (0 %)), and Preussia antarctica (strain CBS 222.89, GenBank KX710224.1; Identities = 492/514 (96 %), eight gaps (1 %)). Closest hits using the LSU sequence are Preussia minimoides (strain MEXU 26355, GenBank KF557659.1; Identities = 886/888 (99 %), no gaps), Sporormiella isomera (strain CBS 166.73, GenBank MH872355.1; Identities = 886 /890 (99 %), two gaps (0 %)), and Sporormiella intermedia (as Preussia intermedia; strain CBS 364.69, GenBank MH878451.1; Identities = 879/ 889 (99 %), one gap (0 %)). No significant hits were obtained when the tef1 sequence was used in blastn and megablast searches. Closest hits using the tub2 sequence had highest similarity to Preussia sp. 10 MP-2020 (strain 18EPLE010, GenBank MT881917.1; Identities = 371/388 (96 %), two gaps (0 %)), Preussia lignicola (strain 18ALIC002, GenBank MT671880.1; Identities = 349 /394 (89 %), 14 gaps (3 %)), and Sporormiella intermedia (strain 18THES003, GenBank MT881987.1; Identities = 346/393 (88 %), 12 gaps (3 %)). Colour illustrations. Mirabib Rock in the Namib Desert, where the sample was collected. Conidiomata on SNA (scale bars = 100 µm); superficial view of conidiomatal wall (scale bar = 50 µm); conidiogenous cells (scale bars = 10 µm); conidia (scale bar = 10 µm). Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl Neriman Yilmaz, Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), Faculty of Natural and Agricultural Sciences, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: neriman.yilmazvisagie@fabi.up.ac.za Don A. Cowan, Centre for Microbial Ecology and Genomics, Department of Biochemistry, Genetics and Microbiology, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: don.cowan@up.ac.za Gillian Maggs-Kölling, Gobabeb-Namib Research Institute, P.O. Box 953, Walvis Bay, Namibia; e-mail: gillian@gobabeb.org © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 322 Persoonia – Volume 45, 2020 Curvularia moringae & Moringomyces phantasmae 323 Fungal Planet description sheets Fungal Planet 1139 & 1140 – 19 December 2020 Curvularia moringae Crous, sp. nov. Etymology. Name refers to the host genus Moringa from which it was isolated. Classification — Pleosporaceae, Pleosporales, Dothideomycetes. Mycelium consisting of pale to medium brown, smooth, branched, septate, 4 – 6 µm diam hyphae. Conidiophores solitary, subcylindrical, erect, geniculate-sinuous, mostly unbranched, 1–10-septate, 20–110 × 5–7 µm, medium brown, smooth. Conidiogenous cells integrated, terminal or intercalary, subcylindrical, geniculate-sinuous to curved or straight, medium brown, smooth, 12 – 20 × 5 –7 µm; hila thickened, darkened, 2 – 4 µm diam. Conidia solitary, arranged in rosettes, ellipsoid, straight to slightly curved, medium brown, finely roughened, 3 – 5-distoseptate, guttulate, apex obtuse, base bluntly rounded with darkened, thickened hilum, 1.5–3 µm diam, (30–)40–48(–51) × (16 –)17– 21(– 23) µm. Culture characteristics — Colonies flat, spreading, with moderate aerial mycelium and smooth, even margin, covering dish after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse iron-grey. Typus. NAmibiA, Gobabeb-Namib Research Institute, on leaves of Moringa ovalifolia (Moringaceae), 19 Nov. 2019, P.W. Crous, HPC 3117 (holotype CBS H-24447, culture ex-type CPC 38873 = CBS 146828, ITS, LSU, gapdh and rpb2 sequences GenBank MW175363.1, MW175403.1, MW173105.1 and MW173117.1, MycoBank MB837854). Notes — Curvularia moringae is phylogenetically distinct from species presently known in the genus (Marin-Felix et al. 2017a, b, 2020). (notes Curvularia moringae continues on Supplementary material page FP1139 & 1140) Moringomyces Crous, gen. nov. Etymology. Name refers to the host genus Moringa from which it was isolated. Classification — Saccotheciaceae, Dothideales, Dothideomycetes. forming intercalary chains of chlamydospores enclosed in mucoid sheath, initially transversely septate, becoming muriformly septate, becoming swollen, eventually forming microsclerotia. Type species. Moringomyces phantasmae Crous. MycoBank MB837855. Mycelium consisting of hyaline, smooth, septate, branched, hyphae. Hyphal cells becoming swollen, brown, roughened, Moringomyces phantasmae Crous, sp. nov. Etymology. Name refers to the host Moringa ovalifolia (Namibian phantom tree), L. phantasma = phantom). Mycelium consisting of hyaline, smooth, septate, branched, 1.5–3 µm diam hyphae. Hyphal cells becoming swollen, brown, roughened, forming intercalary chains of chlamydospores enclosed in mucoid sheath, initially transversely septate, becoming muriformly septate, initially 3 – 4 µm diam, becoming swollen, 8–10 µm diam, eventually forming microsclerotia, more prominent and larger on OA than on SNA, up to 100 µm diam. Culture characteristics — Colonies flat, spreading, surface folded, with sparse aerial mycelium and feathery, lobate margin, reaching 45 mm diam after 2 wk at 25 °C. On MEA, PDA and OA surface and reverse iron-grey. Typus. NAmibiA, Gobabeb-Namib Research Institute, on flower of Moringa ovalifolia (Moringaceae), 19 Nov. 2019, P.W. Crous, HPC 3130 (holotype CBS H-24449, culture ex-type CPC 38883 = CBS 146830, ITS and LSU sequences GenBank MW175364.1 and MW175404.1, MycoBank MB837856). Notes — Moringomyces is related to the genera Arxiella, Aureobasidium and Pseudosydowia. Other than the microsclerotia, Moringomyces did not form any conidiomata or conidia in culture, making morphological comparisons difficult. Genera in this complex have similar culture characteristics, namely pigmented hyphae that are strongly constricted at septa, encased in mucilage, and aggregations of hyphal cells that tend to form microsclerotia. (notes Moringomyces phantasmae continues on Supplementary material page FP1139 & 1140) Colour illustrations. Moringa ovalifolia tree growing in the Namib Desert. Left column Curvularia moringae. Conidiophores and conidiogenous cells giving rise to conidia. Right column Moringomyces phantasmae. Colonies on SNA; hyphae encased in mucilage; microsclerotium. Scale bars = 10 µm. Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl Neriman Yilmaz, Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), Faculty of Natural and Agricultural Sciences, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: neriman.yilmazvisagie@fabi.up.ac.za Don A. Cowan, Centre for Microbial Ecology and Genomics, Department of Biochemistry, Genetics and Microbiology, University of Pretoria, Private Bag X20, Hatfield 0028, Pretoria, South Africa; e-mail: don.cowan@up.ac.za Gillian Maggs-Kölling, Gobabeb-Namib Research Institute, P.O. Box 953, Walvis Bay, Namibia; e-mail: gillian@gobabeb.org © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 324 Persoonia – Volume 45, 2020 Neodothiora populina 325 Fungal Planet description sheets Fungal Planet 1141 – 19 December 2020 Neodothiora Crous, G.C. Adams & Winton, gen. nov. Etymology. Name refers to its superficial resemblance of the genus Dothiora. Classification — Dothioraceae, Dothideales, Dothideomycetes. Conidiomata solitary, erumpent, brown, subglobose, pycnidial, with central ostiole, exuding a crystalline mucoid conidial cirrhus; wall of 6 – 8 layers of brown textura angularis. Conidiophores reduced to conidiogenous cells lining the inner cavity, hyaline, smooth, ampulliform, proliferating percurrently. Conidiogenous cells also occurring solitary on superficial hyphae, subcylindrical, hyaline, smooth, proliferating percurrently at apex. Conidia solitary, hyaline, smooth, aseptate, guttulate, ellipsoid, apex subobtuse, tapering to truncate apex. Type species. Neodothiora populina Crous, G.C. Adams & Winton. MycoBank MB837857. Neodothiora populina Crous, G.C. Adams & Winton, sp. nov. Etymology. Name refers to the host genus Populus from which it was isolated. bark around points of inoculation, which resembled those that developed in culture (on agar and on PNA). On PNA: Conidiomata solitary, erumpent, brown, subglobose, pycnidial, with central ostiole, 130 –180 µm diam, exuding a crystalline mucoid conidial cirrhus; wall of 6 – 8 layers of brown textura angularis. Conidiophores reduced to conidiogenous cells lining the inner cavity, hyaline, smooth, ampulliform, proliferating percurrently, 5 –7 × 4 – 6 µm. Conidiogenous cells also occurring solitary on superficial hyphae, ampulliform to subcylindrical, hyaline, smooth, 5 –10 × 2 – 5 µm, proliferating percurrently at apex. Conidia solitary, hyaline, smooth, aseptate, guttulate, ellipsoid, apex subobtuse, tapering to truncate apex, 5 – 6(–7) × 2.5– 3 µm. Culture characteristics — Colonies flat, spreading, with sparse aerial mycelium and feathery, uneven margin, covering dish after 2 wk at 25 °C. On MEA surface mucoid, saffron, reverse saffron with patches of umber; on PDA surface and reverse umber, margin black; on OA surface umber. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to ‘Uncultured fungus’ (strain UPSC_A12_12, GenBank GU564975.1; Identities = 525 /527 (99 %), one gap (0 %)), Scleroconidioma sphagnicola (strain JJ-18-24, GenBank MK880096.1; Identities = 563/591 (95 %), 13 gaps (2 %)), Rhizosphaera macrospora (strain ARSL_071114.1, GenBank Typus. usA, Alaska, -148.7762872 64.63940972, on stem cankers of Populus tremuloides (Salicaceae), 24 June 2018, G. Adams & L.M. Winton (holotype CBS H-24556, culture ex-type CPC 39399 = CBS 147087, ITS, LSU, tef1 and tub2 sequences GenBank MW175365.1, MW175405.1, MW173127.1 and MW173142.1, MycoBank MB837858). Additional materials examined. usA, Alaska, -148.7672229 64.64259316, on stems of P. tremuloides, 25 June 2018, G. Adams & L.M. Winton, CPC 39397 = CBS 147085, ITS sequence GenBank MW175366.1; usA, Alaska, -148.3288146 64.73376442, on stems of P. tremuloides, 19 June 2018, G. Adams & L.M. Winton, CPC 39398 = CBS 147086, ITS sequence GenBank MW175367.1; Alaska, Bonanza Creek Experimental Forest, -148.33106. 64.73243, on stems of P. tremuloides, 15. Sept. 2020, L.M. Winton, Univ. of Alaska Herbarium (ALA) H1280665, H1280672; ibid., on stems of P. tremuloides, Jan. 2020, L.M. Winton, H1280666–H1280671. Notes — Neodothiora is reminiscent of the genus Dothiora, having Dothichiza and hormonema-like morphs in culture (Crous & Groenewald 2016, 2017). However, it clusters apart from the type species, D. pyrenophora, and thus a new genus is herewith introduced to accommodate this pathogen, which is associated with severe cankers of Populus tremuloides in Alaska. In field inoculations, conidiomata developed on the tree Colour illustrations. Stem canker on Populus tremuloides. Conidioma on PNA; conidiomata on SNA; conidiogenous cells; conidiogenous cells giving rise to conidia; conidia. Scale bars: conidiomata = 150 µm, all others = 10 µm. (text continues on Supplementary material page FP1141) Supplementary material FP1141-1 The first of 1 000 equally most parsimonious trees obtained from the LSU alignment (57 sequences including the outgroup; 809 characters including alignment gaps analysed: 566 constant, 122 variable and parsimony-uninformative and 121 parsimony-informative) using PAUP* v. 4.0b10 (Swofford 2003). Tree statistics: TL = 496, CI = 0.653, RI = 0.813, RC = 0.531. Parsimony bootstrap support values > 74 % and Bayesian posterior probabilities (PP) > 0.79 are shown at the nodes and thickened lines represent branches present in the parsimony strict consensus tree. The Bayesian analysis using MrBayes v. 3.2.7a (Ronquist et al. 2012) resulted in a Bayesian consensus phylogram based on 633 002 sampled trees and 180 unique site patterns (data not shown). The scale bar represents the number of changes. The taxonomic novelties described in this study are highlighted with bold text and coloured blocks. GenBank accession and culture/specimen numbers are indicated behind the species names. The two orders are indicated to the left of the tree at the basal branches. The tree was rooted to Diaporthe perjuncta (GenBank NG_059064.1). The alignment and tree were deposited in TreeBASE (Submission ID 27179). FP1141-2 The first of 414 equally most parsimonious trees obtained from the ITS alignment (50 sequences including the outgroup; 518 characters including alignment gaps analysed: 332 constant, 74 variable and parsimonyuninformative and 112 parsimony-informative) using PAUP* v. 4.0b10 (Swofford 2003). Tree statistics: TL = 377, CI = 0.647 RI = 0.882, RC = 0.571. Parsimony bootstrap support values > 74 % and Bayesian posterior probabilities (PP) > 0.79 are shown at the nodes and thickened lines represent branches present in the parsimony strict consensus tree. The Bayesian analysis using MrBayes v. 3.2.7a (Ronquist et al. 2012) resulted in a Bayesian consensus phylogram based on 161 252 sampled trees and 176 unique site patterns (data not shown). The scale bar represents the number of changes. The taxonomic novelties described in this study is highlighted with bold text and coloured blocks. GenBank accession and culture /specimen numbers are indicated behind the species names. The tree was rooted to Dothidea sambuci (GenBank NR_111220.1). The alignment and tree were deposited in TreeBASE (Submission ID 27179). Pedro W. Crous & Johannes Z. Groenewald, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: p.crous@wi.knaw.nl & e.groenewald@wi.knaw.nl Gerard C. Adams, Department of Plant Pathology, 406D Plant Science Hall, 1875 N. 38th Street, University of Nebraska, Lincoln, NE, USA; e-mail: gadams3@unl.edu Loretta M. Winton, U.S.D.A. Forest Service, Forest Health Protection, 3700 Airport Way, Fairbanks, AK 99709, USA; e-mail: loretta.winton@usda.gov © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 326 Persoonia – Volume 45, 2020 Amanita domingensis 327 Fungal Planet description sheets Fungal Planet 1142 – 19 December 2020 Amanita domingensis Angelini & Vizzini, sp. nov. Etymology. Referring to the place of the first collection, Santo Domingo, the capital of the Dominican Republic. Classification — Amanitaceae, Agaricales, Agaricomycetes. Pileus (4.5–)5.5–6(–8.5) cm diam, campanulate, then convex, plane convex, sometimes depressed at the centre at maturity and then with a poor developed obtuse umbo, margin striated up to 1/3 of the radius; surface glabrous, opaque, oily, viscid when moist, ash grey, covered with general veil remnants in the form of whitish grey floccose patches or warts, more abundant on the centre. Lamellae free, sometimes distant, straight, interspersed with lamellulae of varying length, 0.5 cm wide, white then white-yellow and with a finely eroded grey edge. Stipe 8 –10 × 0.7– 0.9 cm, cylindrical, straight or slightly sinuous, narrowing and flared upwards, internally fistulous; surface covered with small, grey fibrillose squamules, sometimes becoming progressively more snakeskin-patterned and greyer towards the base, on a white background. Volva ± membranous, slightly adherent, low, internally white, externally whitish in the hypogeous part, grey in the emerging part, tending to fracture horizontally forming one or more rings of dark volval material at the stipe base. Annulus absent. Context thin, 0.2 – 0.3 cm thick (in the pileus), white. Odour and taste not distinctive. Spores (10 –)11–12(–13.5) × 8 – 9 µm (av. 11.4 × 8.5 µm, Qm = 1.35), broadly ellipsoid to oblong, thin-walled, mostly containing one large drop, hyaline, inamyloid, with slightly prominent and eccentric apiculus. Basidia 30 – 50 × 13 –15 µm, clavate, mostly tetrasporic, sometimes bisporic, with sterigmata up to 5 µm long. Marginal cells present but not abundant, not emerging over the basidia, mostly consisting of single thin-walled elements, sometimes bi-catenulate, rarely tri-catenulate, whose basal element, when present, has a mostly ovoid shape, while the terminal one has a pyriform or ovoid-claviform shape, 16 – 22 × 10 –11 µm wide. Partial veil consisting of sphaeropedunculate elements on the lamellar edge, 25 – 30 × 22 – 25 µm wide. Universal veil (volva) mixed in structure (membranous + spherocytic) consisting of intertwining hyphae 5 – 6 µm wide, with also subglobose to sphaeropedunculate elements, 30–45 × 20 – 35 µm wide. Subhymenium a puzzle layer of cubic-multifaceted cells, about 50 µm wide. Lamellar trama divergent. Pileipellis an ixocutis of stretched and variously intertwined hyphae, with rounded terminals up to 7 µm wide, completely immersed in a hyaline gelatinous layer. Context of non-inflated hyphae, 3 – 9 µm wide. Stipitipellis a cutis, similar to pileipellis, but non-gelatinized, consisting of parallel non-inflated hyphae, 5–7 µm wide, covered by a layer of hyphae of the universal veil with elongated, pyriform, large terminal elements, 70–180 µm wide; occasionally, with pedunculate spherocytes, residues of the partial veil, similar in shape and size to those of the lamellar edge, 25 – 35 × 20 – 25 µm wide. Stipititrama acrophysalidic. Clamp-connections absent. Habitat & Distribution — Exclusive in deciduous woods (probably associated with Coccoloba diversifolia), from the plains (but far from the beaches) to the hills, gregarious or as single specimens, in autumn and winter. Common. Typus. DomiNicAN republic, National Garden of Santo Domingo, Distrito Nacional, six specimens collected on litter of deciduous wood, 24 Nov. 2014, C. Angelini (holotype JBSD130784, ITS and LSU sequences GenBank MT991052 and MT991057, MycoBank MB837379). Additional materials examined. DomiNicAN republic, National Garden of Santo Domingo, Distrito Nacional, on litter of deciduous wood, 18 Nov. 2013, C. Angelini JBSD130785 (ITS and LSU sequences GenBank MT991053 and MT991058); Puerto Plata, Sosua, 25 Dec. 2016, C. Angelini JBSD130786 (ITS and LSU sequences GenBank MT991054 and MT991059). Notes — Amanita domingensis is one of the few Dominican Amanita species not in association with conifers and represents the most common and abundant Amanita in the deciduous forests. It belongs in sect. Vaginatae (subg. Amanita) where, in the molecular analysis, it occupies an isolated position. Amanita arenicola from Puerto Rico and the British Virgin Islands, which is common on the beaches in association with Coccoloba uvifera, is distinguished from the new species by its exclusively sabulicolous habitat, the veil, the stipe, the lamellae and the lamellar edge that are completely white at all developmental stages (Miller et al. 2000) and the different ITS sequence, 448/527 bp (85 %) similar. Amanita antillana described from Trinidad and Tobago as an ectomycorrhizal associate of Coccoloba pubescens and Haematoxylum campechianum, differs by the olive brown pileus, usually devoid of velar remnants and with a shortly striated margin, the fragile ochraceous brown volva that often disappears from the stipe at maturity, and broader spores, 10 –13.5(–15) × 7.5 –11.5(–13) µm (Dennis 1952, Pegler 1983). Supplementary material Colour illustrations. Dominican Republic, Puerto Plata, Sosua, deciduous natural forest. Fresh basidiomes in field (holotype JBSD130784); volva detail; fresh basidiomes in field (JBSD130785); spores. Scale bars = 1 cm (basidiomes), 10 μm (spores). FP1142 Maximum-likelihood analysis of the nrITS region of Amanita sect. Vaginatae species was performed with RAxML v. 8 (Stamatakis 2014) using the GTR+G model (1 000 bootstrap replicates, bootstrap support values ≥ 70 % are shown). The scale bar represents the number of nucleotide changes per site. Alfredo Vizzini & Francesco Dovana, Department of Life Sciences and Systems Biology, University of Torino, Viale P.A. Mattioli 25, I-10125 Torino, Italy; e-mail: alfredo.vizzini@unito.it & francescodovana@gmail.com Claudio Angelini, Herbario Jardín Botánico Nacional Dr. Rafael Ma. Moscoso, Santo Domingo, Dominican Republic and Via Cappuccini, 78/8 – 33170 Pordenone, Italy; e-mail: claudio_angelini@libero.it © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 328 Persoonia – Volume 45, 2020 Austroboletus asper 329 Fungal Planet description sheets Fungal Planet 1143 – 19 December 2020 Austroboletus asper K. Syme, Bonito, T. Lebel, Fechner & Halling, sp. nov. Classification — Boletaceae, Boletales, Agaricomycetes. Typus. AustrAliA, Western Australia, Denmark, 1874 South Coast Highway, SE section, S34.9861° E117.2876°, 5 May 2013, K. Syme 2828 (holotype MEL2371703, isotype NY02072470, ITS and rpb1 sequences GenBank KP242152 and KP242055, MycoBank MB836723). Pileus 3.5–9.5(–22) cm broad when expanded, conico-convex to convex to plano-convex, rarely glutinous to viscid when young and fresh, becoming minutely areolate tomentose to matted tomentose and dry as gluten dissipates, often with a suede-like texture, with white, sterile appendiculate veil remnants, cinnamon brown to brown to cocoa brown (8A3,2; 7E8,7,6; 7D7,6; 6B3; 6C–D–E8,7,6,5; 5C6; Kornerup & Wanscher, 1983). Flesh white, unchanging, up to 2 cm deep, with mild odour and taste. Tubes depressed to deeply depressed around stipe, white when young, soon pinkish brown (8D4; 7C4), usually staining brown (6E8,7). Stipe 3.5 – 8(–12) cm long, 0.6 –1.7(– 3.4) cm broad, equal to subclavate, strict or curved, sometimes slightly tapered at base, alveolate to lacunose-reticulate, sometimes coarsely so, white above, occasionally pale yellowish, sometimes pale brownish below on reticulum when fresh, often yellow to brownish orange (5B7, 6C7) at base, especially with age, rarely viscid below when young and moist, otherwise dry, with interior white, unchanging, pithy to hollow with age, white or sometimes pale yellow in the base, with white rhizomorphs. Additional material examined. AustrAliA, Queensland, Tablelands, Davies Creek National Park, Davies Creek Road, 12.6 km from Kennedy Hwy, S17.0264° E145.6°, 680 m, 19 Feb. 1992, Halling 6827 (PERTH, NY, LSU sequence GenBank KP242247); Wide Bay District, Great Sandy National Park, Fraser Island, near Lake Birrabeen, S25.4918° E153.054°, 108 m, 4 June 2009, Halling 9159 (BRI, NY); Fraser Island, S25.4095° E153.086°, 112 m, 6 June 2009, Halling 9172 (BRI, NY, LSU, rpb1 and rpb2 GenBank sequences MT921383, MT932084 and MT928122); Fraser Island, Cornwells Road near Kingfisher Bay, S25.4019° E153.031°, 94 m, 24 May 2010, Halling 9362 (BRI, NY, ITS, rpb1 and rpb2 sequences GenBank KP242158, KP242047 and KP242087); Fraser Island, Northern Road, ± 1 km N of Cornwells Road, S25.4253° E153.059°, 100 m, 26 May 2010, Halling 9393 (BRI, NY, rpb1 sequence GenBank KP242057); Victoria, Colac Otway, Carlisle Sate Park, Cricket Pitch Track, 7.5 km W of Gellibrand, S38.5428° E143.477°, 170 m, 9 May 2005, Halling 8685 (MEL, NY, ITS sequence GenBank KP242166); Western Australia, Denmark, off Sunny Glen Rd, state forest N of Plantagenet loc 6722, 6 June 1996, Syme 877 (PERTH, ITS and rpb2 sequences GenBank KP242218 and KP242111); adjoining Plantagenet loc 6721, S34.9315° E117.4435°, 6 May 1998, Syme 955 (PERTH); Walpole-Nornalup National Park, The Knoll lower walk, S34.9938° E116.727°, 18 May 2001, Syme 1139 (PERTH, MEL, LSU and ITS sequences GenBank KP242267 and KP242165). Spores (12.1–)14.3 –18.7(– 20) × 4.4 – 5.5 µm (av. = 16.23 × 4.92 μm, Q = 3.33, spores n = 159, specimens n = 8), faintly wrinkled to very finely and uniformly rugulose to irregularly granular (light microscope) or irregularly foveate with low, short meandering ridges and low, isolated tubercles (SEM), weakly dextrinoid in Melzer’s. Basidia 29–36 × 6–11 μm, four-sterigmata, clavate, hyaline. Tube trama boletoid and divergent, inamyloid, with occasional laticiferous elements, with hyphae, 3.5–10 μm broad. Hymenial cystidia 50 – 60 × 12 – 20 μm, scattered, embedded in the hymenium with rostrate portion protruding, fusoid rostrate to ventricose rostrate with septum separating a broad rostrum from ventricose portion, hyaline, thin-walled, inamyloid. Pileus trama inamyloid, hyaline in KOH, with hyphae 5 – 8 μm broad. Pileipellis a tangled, erect, dense trichodermium, soon collapsing; hyphae with rare hyaline encrustations, hyaline or otherwise a pale ochraceous in KOH, 5–8 μm broad. Stipitipellis a tangled trichodermium of thin-walled, hyaline, soon collapsing hyphae, rarely with obvious end cells. Habit, Habitat & Distribution — Solitary to gregarious on soil or sand with Agonis flexuosa, Allocasuarina fraseriana, A. littoralis, Corymbia calophylla, Eucalyptus diversicolor, E. guilfoylei, E. marginata, E. pilularis, E. racemosa, Leptospermum sp., Lophostemon sp., and Syncarpia glomulifera. At present, known in Queensland, Tasmania, Victoria, Western Australia. Additional GenBank sequences. rpb1: KP242085; rpb2: KP242126, KP242127; ITS: KP242164, KP242173, KP242174, KP242186, KP242187, KP242204, KP242216 Etymology. Asper (rough), in reference to the ornamentation of the basidiospores. Notes — In Australia, there are other Austroboletus spp. with similar macromorphology, i.e., pileus colour, texture, and viscidity when fresh and moist. Degrees of separation are based on geographic distribution along with spore size and ornamentation. Austroboletus occidentalis appears to be restricted to Western Australia and has shorter, more citriform spores with similar, but coarser ornamentation and possesses a smooth plage (e.g., Syme 2082: PERTH8105421, NY02449690), a feature not noted in the protologue (Watling & Gregory 1986). Two other currently undescribed species (Austroboletus sp. 5, sp. 6) are only known from Queensland and tropical Northern Territory. A key feature of these latter is the asperulate spore ornamentation that can be difficult to see with a compound light microscope. However, with patience, high resolution optics equipped with Nomarski DIC lenses, a discrete ornamentation can be observed especially if the plage can be seen in adaxial and profile views. Colour illustrations. Sclerophyll forest of Eucalyptus diversicolor, Corymbia calophylla and Agonis flexuosa near Denmark, Western Australia (photo K. Syme). Habit (Syme 2828, type); spores with DIC light microscope and with SEM; pileipellis; hymenial cystidium. Scale bars = 5 cm (habit), 5 µm (SEM), 10 µm (light micrographs). Katrina Syme, National Herbarium of Victoria, Royal Botanic Gardens Victoria, South Yarra, Victoria 3141, Australia; e-mail: katrinasyme@gmail.com Gregory Bonito, Department of Plant Soil and Microbial Sciences, 1066 Bogue Street, Michigan State University, East Lansing MI, 48824 USA; e-mail: bonito@mail.msu.edu Teresa Lebel, Botanic Gardens & State Herbarium, Adelaide, South Australia, Australia; e-mail: teresa.lebel@sa.gov.au Nigel Fechner, Queensland Herbarium, Mt Coot-tha Road, Toowong, Brisbane, Queensland 4066, Australia; e-mail: nigel.fechner@des.qld.gov.au Roy E. Halling, Inst. Systematic Botany, New York Botanical Garden, 2900 Southern Blvd, Bronx, NY, USA 10458-5126; e-mail: rhalling@nybg.org © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 330 Persoonia – Volume 45, 2020 Cantharellus betularum 331 Fungal Planet description sheets Fungal Planet 1144 – 19 December 2020 Cantharellus betularum Voitk & Thorn, sp. nov. Etymology. Betularum (Latin: of birches) refers to the tree associates of the species. Classification — Hydnaceae, Cantharellales, Agaricomycetes. Pileus 20–70 mm diam, margins inrolled, becoming plane, then funnel-shaped and irregularly wavy, opaque, yellow-gold, with thin amethyst coating that breaks up into small scales, becoming violet brown, then brown; scales often absent and amethyst colour not common. Hymenium folds moderately spaced, wide, blunt, sinuous, forked, cross-veined and anastomosing, deeply decurrent to almost absent; pale yellow to almost white. Stipe 5 – 25 × 30 – 65 mm, enlarging upwards, solid, yellow. Context whitish yellow; odour sweet and fruity. All tissues stain reddish brown with injury or prolonged exposure. Aberrant forms in exposed habitats vary from solitary pegs to fused multicephalic basidiomes. Basidiospores (two observers, five collections, seven sporocarps, 131 spores) (7.7–)8.7–14.3 × (3.9 –)4.6 – 7.1(–7.7) µm (av. 10.4 × 5.6 µm), av. Q = 1.9; elliptical-oblong, usually narrower at the apex, slightly bent, with an asymmetric constriction; content homogeneous. Basidia 65–90 × 7.7–11.6 µm; 4 – 6-spored; clavate. No cystidia. Clamp connections in all tissues. Wide range in micromorphology between individual basidiomes and collections. Habitat & Distribution — Solitary or gregarious in leaf litter of Betula, hitherto known from three sites in the Bay of Islands region of western Newfoundland. 48.99, -057.76, 30 Aug. 2011, A. Voitk, 11.08.30.av02 (DAOM 734023), LSU: KX592688, KX592689; same locale, 21 Aug. 2013, M. Voitk, 13.08.21.av01 (DAOM 734026), LSU: KX592698, KX592699; near Frenchman’s Cove, 49.046, -058.185, E. Humber, 14.09.01.av01 (DAOM 734020). Notes — Using LSU sequence data, we previously reported this species as Cantharellus amethysteus (Thorn et al. 2017), but ITS, LSU and tef1 are all required to differentiate these two taxa phylogenetically. Cantharellus betularum differs from C. amethysteus by growing on a different continent (impediment to continued genetic mixing), in a region about 10 °C colder, on average, and with birch, not the oak (Quercus) or beech (Fagus; both Fagaceae and not native to Newfoundland) most commonly recorded with C. amethysteus. We have not seen C. amethysteus, but the description by Buyck (2000) suggests that its amethyst scales are more consistent and prominent than those of C. betularum, which are often absent, and its spores are broader (av. 6.5 vs 5.6 µm). Other North American vinaceousviolaceous species are not found in Newfoundland (Buyck & Hofstetter 2011, Herrera et al. 2018), and C. betularum has not been reported outside the Island. Amethyst scales, longer spores, association with birch, and sequence data separate it from C. camphoratus and C. enelensis, the two other golden chanterelles in Newfoundland (Thorn et al. 2017). Typus. CANADA, Newfoundland and Labrador, Humber Village, trail to Barry’s Lookout, 48.988, -057.792, 159 m a.s.l., in leaf litter under Betula papyrifera, B. cordifolia and B. alleghaniensis (Betulaceae), 14 Sept. 2013, Andrus Voitk 13.09.14.av01 (holotype DAOM 721702, isotype DAOM 734027, nrLSU sequences GenBank KX592700–KX592701, MycoBank MB836965). Collection and sequence data of 14 paratypes: All same site as holotype (CANADA, Newfoundland and Labrador, Humber Village, trail to Barry’s Lookout, 48.988, -057.792, 159 m a.s.l.) and collector (A. Voitk) except where noted below, 24 Aug. 2008, 09.08.24.av04 (DAOM 72173), Tef1: MN181459, ITS-LSU: MN206942; 25 Aug. 2010, 10.08.25.av02 (DAOM 734021), Tef1: MN181461, ITS-LSU: MN206944, MN206945; 12 Aug. 2011, 11.08.12.av01 (DAOM 734022), Tef1: MN181460, ITS-LSU: MN206943; 10 Aug. 2012, 12.08.10.av01 (DAOM 734016); 2 Sept. 2012, 12.09.02.av11 (DAOM 734024), Tef1: KX592690, LSU: KX592691, KX592692; 3 Oct. 2012, 12.10.03.av01 (DAOM 734025), Tef1: KX592693, ITS: KX592696, KX592697, LSU: KX592694, KX592695; 11 Aug. 2013, 13.08.11.av01 (DAOM 734017); 30 Sept. 2013, 13.09.30.av04 (DAOM 734018), 1 Oct. 2013, 13.10.01.av02 (DAOM 734019), 30 Sept. 2017, M. Voitk, 17.09.30.av01 (DAOM 984767), Tef1: MN181458, ITS: MN206939, LSU: MN206940, MN206941; 2 Sept. 2018, 18.09.02.av01 (DAOM 984768); Humber Village, trail to Weldon’s, Colour illustrations. Canada, Newfoundland, Humber Village, near trail to Barry’s Lookout, a mixed forest dominated by Betula papyrifera, B. cordifolia and B. alleghaniensis, where the holotype was collected. Left: typical appearance of C. betularum, with stipitate basidiomata, one fused multicephalic specimen and two peg-like specimens. Note the lighter, blunted hymenial folds, almost absent on the peg form, and brownish orange staining. Centre: close-up of pileus, showing lavender scales, when present. Right: basidiospores, original magnification × 1 000 (modified from Thorn et al. 2017: f. 2B, with permission). Scale bars = 1 cm (basidiomes) and 10 μm (basidiospores). One of 119 equally most parsimonious trees based on sequences of nrITS, nrLSU, and tef1, with node support above branches from Bayesian inference (BI: MrBayes v. 3.2.6, Ronquist et al. 2012), 1 000 bootstrap replicates in maximum likelihood (ML; MEGA X, Kumar et al. 2018, Stecher et al. 2020), and 100 bootstrap replicates in maximum parsimony (MP; PAUP v. 4.0b10, Swofford 2003), and percent consensus among the 119 equally most parsimonious MP trees below. Branches with less than 50 % support are marked with dashes (--) and those that collapsed in a particular analysis are marked with asterisks (**). New sequences are indicated in bold, and sequences from types are indicated as HT (holotype) or NT (neotype). R. Greg Thorn & Alicia Banwell, Department of Biology, University of Western Ontario, London, Ontario, N6A 5B7, Canada; e-mail: rgthorn@uwo.ca & abanwel2@uwo.ca Jee In Kim, Faculty of Computer Science, Dalhousie University, Halifax, Nova Scotia, B3H 4R2, Canada; e-mail: jeein.j.kim@gmail.com Renée Lebeuf, 775, Rang du Rapide Nord, Saint-Casimir, Québec, G0A 3L0, Canada; e-mail: renee.lebeuf@gmail.com Andrus Voitk, 13 Maple St, Humber Village, Newfoundland and Labrador, A2H 2N2, Canada; e-mail: seened@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 332 Persoonia – Volume 45, 2020 Chaetothyrina spondiadis 333 Fungal Planet description sheets Fungal Planet 1145 – 19 December 2020 Chaetothyrina spondiadis Fuentes-Aponte, K. Kim & Romberg, sp. nov. Etymology. Named for Spondias, the host genus from which this fungus was collected. Classification — Phaeothecoidiellaceae, Mycosphaerellales, Dothideomycetes. Causes flyspeck on fruits of Spondias. Ascomata thyrothecial, circular, medium to dark brown, gregarious to solitary, superficial, 164.5 – 254 µm diam, ostiolate, margin entire to slightly irregular. Setae 49 –112.5 µm long, wider at the base, scattered on the surface of the thyrothecia, straight, unbranched, septate, brown, smooth, easily removed. Upper wall consisting of 2 – 3 layers of cells, dark brown, textura epidermoidea. Hamathecium consisting of septate, hyaline pseudoparaphyses, 1.5 – 2 µm wide, sometimes branched at the tip. Asci bitunicate, oblong to pyriform, 24.7–50.8 × 8.8–16.4 µm, with eight biseriate ascospores. Ascospores hyaline, ovoid to elongated ovoid, 11.3 –15.75 × 2.5 – 5.6 µm, 1-septate, often slightly constricted at septum, ends rounded, walls smooth. Culture characteristics — Colonies slow-growing, reaching 15–30 mm diam after 35 d at 25 °C on MEA. Colony pulvinate, circular, entire, with a light grey surface, and reverse dark irongrey. Typus. USA, Puerto Rico, Hatillo, on fruits of Spondias mombin (Anacardiaceae), Nov. 2018, S. Fuentes-Aponte (holotype BPI 911218, culture ex-type CBS 145915, ITS and LSU, sequences GenBank MT339448 and MT339447, MycoBank MB835259). Additional material examined. USA, Puerto Rico, San Juan, S. mombin, 1961, M. Farr, BPI 644792; San Juan, Spondias sp., 1966, F. Pollack, BPI 644782; San Juan, Spondias sp., 1969, F. Pollack, BPI 646405; San Juan, S. dulcis, 1970, F. Pollack, BPI 646519. – Intercepted specimens: USA, intercepted in Miami, Florida, entering from Jamaica, S. cytherea, 1963, F. Pollack, BPI 646407; intercepted in New York, New York, entering from Brazil, S. mombin, 1964, F. Pollack, BPI 646446; entering from Trinidad, S. dulcis, 1966, A. Watson, BPI 646243. Notes — Chaetothyrina was described in 1913 by Theissen, with type species Chaetothyrina musarum. Several species have been described in the genus, mainly on tropical hosts including Artocarpus (Chaetothyrina artocarpi), Mangifera (Chaetothyrina guttulata), Anacardium (Chaetothyrina megalospora) and Musa (Chaetothyrina musarum) (Singtripop et al. 2016). The measurements of salient characters for most of these species overlap. Stevenson (1975) identified the fungus causing flyspeck on Spondias cytherea and Spondias mombin in Puerto Rico as Chaetopeltopsis tenuissima which was later transferred to Chaetothyrina as C. tenuissima (Müller & Von Arx 1962). Chaetothyrina tenuissima (Asterina tenuissima) was described from Hevea brasilensis in Sri Lanka by Petch (1906) with the following characters: ‘perithecia 130 –160 µm diam, asci 30 – 40 × 9 –12, spores 13 × 4, one-septate, constricted, fusoid, hyaline’. Other superficial, thyrothecial fungi with similar ascospores reported from Spondias include Stomiopeltis sp. reported from Venezuela on Spondias mombin and Schizothyrium sp. reported from Spondias purpurea in the West Indies. Chaetothyrina spondiadis is genetically distinct from both Stomiopeltis and Schizothyrium and clearly belongs to Chaetothyrina. It is morphologically distinct from C. tenuissima, having larger thyrothecia, asci and ascospores. Few described Chaetothyrina species have sequences available in public databases. In a megablast search of the NCBI GenBank, ITS sequences of C. spondiadis showed highest identity to Chaetothyrina guttulata (GenBank NR_153923.1, 98.17 %) and Chaetothyrina musarum (GenBank KX372275.1, 96.88 %). Alignment of the ITS regions of these and other fungi in the Capnodiales revealed several indels and SNPs between the sequences of the two other species of Chaetothyrina available publicly and Chaetothyrina spondiadis. Colour illustrations. Puerto Rico, type locality, tree of Spondias mombin. Thyrothecia on host; thyrothecium; asci; ascospores. Scale bars = 10 µm. Megan Romberg, Katherine Kim & Stephanie Fuentes-Aponte, USDA APHIS PPQ NIS, 10300, Baltimore Avenue, Beltsville, MD 20705, USA; e-mail: Megan.k.romberg@usda.gov, Katherine.s.kim@usda.gov & Stephanie.e.fuentes@usda.gov © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 334 Persoonia – Volume 45, 2020 Circinella lampensis 335 Fungal Planet description sheets Fungal Planet 1146 – 19 December 2020 Circinella lampensis E. Alvarez, C. Muñoz & I. Fernandez, sp. nov. Etymology. Referring to Lampa, where this fungus was collected, Lampa Caves, Santiago, Chile. Classification — Lichtheimiaceae, Mucorales, Mucoromycetes. Hyphae hyaline, 5 –10 μm wide, thin- to thick-walled, smooth, aseptate. Sporangiophores erect, 2 – 20 mm high, branched, hyaline to brownish when older, producing sporangia mainly in umbels of 4 – 6, circinate branches, often uniseptate stalks; sporangia spherical or subglobose, brown to black in transmitted light, 40 –75 µm diam, but mostly about 55 µm; columellae ranging from 15 – 30 µm, but usually 20 µm, globose to subglobose or pyriform in shape; sporangiospores (4.5 –)5 –7.5 µm diam, mostly 6 µm, globose to subglobose, biconcave in the frontal view, singly hyaline to slightly coloured. Zygospores and chlamydospores not observed. Culture characteristics — Colonies on potato dextrose agar (PDA) attaining 90 mm diam after 9 –10 d at 25 °C, cottony, whitish to light greyish, reverse hyaline. Growth observed at 15 and 25 °C, but no growth at 5 and 37 °C. Typus. chile, Santiago, Lampa caves, from soil, Jan. 2020, E. Alvarez, C. Muñoz & I. Fernandez (holotype ChFC-564 in Chilean Fungal Collection preserved in a metabolically inactive state, ex-type culture ChFC-2020564; ITS and LSU sequences GenBank MT764259 and MW082021, MycoBank MB836221). Notes — Based on BLAST search results, the closest hits with the ITS sequence were Circinella umbellata (GenBank JN205858; Identities = 596/611 (97.55 %), six gaps (0 %)) and C. minor (GenBank MH854640; Identities = 588/611 (96 %), eight gaps (1 %)). Phylogenetic inference, performed using the ITS sequences of different Circinella spp., including the type species C. umbellata, demonstrated that our fungus represents a new species of the genus Circinella, being closely related to the species C. umbellata. Macroscopically, C. lampensis resembles C. umbellata (Hesseltine & Fennell 1955). Both species showed whitish greyish colonies on all media tested. However, microscopically, C. lampensis presents umbels of up to six sporangia, contrasting to C. umbellata which produce umbels of up to 12 sporangia. Also, C. lampensis differs from C. umbellata in having smaller sporangia (up to 75 µm diam vs up to 120 μm diam in C. umbellata), usually smaller sporangiospores (5 –7.5 μm vs 4.5 –10.5 μm in C. umbellata), and smaller columellae (15 – 30 μm vs 84 – 90 μm in C. umbellata). In addition, C. minor can be distinguished from C. lampensis due the larger size of its sporangia and columellae (40–90 μm, and 12–75 μm vs 40–75 μm, and 15 – 30 μm, respectively). Circinella ramosa CGMCC:3.14094 Circinella ramosa CGMCC:3.14091 T Circinella chinensis CBS 140.28 T Circinella chinensis FMR 15770 Circinella muscae NBRC4457 Circinella muscae CBS 107.13 93 Circinella nodulosa CGMCC 3.14102 T 90 Circinella mucoroides CYD1000719 Circinella angarensis CBS 172.62 86 Circinella angarensis CBS 173.62 NT 100 Circinella minor NRRL 1353 Circinella minor CBS 142.81 95 Circinella lampensis CHFC564 T 96 95 Circinella umbellata NRRL1713 Circinella umbellata CBS 101.16 Mucor simplex CBS 428.80 0.05 Colour illustrations. Lampa caves, Santiago de Chile; colony after 7 d at 25 °C on PDA; umbel with sporangia; sporangia and columella; sporangiospores. Scale bars = 50 µm (sporangia borne in umbel), 10 µm (all others). Maximum Likelihood tree obtained from ITS sequences of our isolate and sequences retrieved from the GenBank nucleotide database. The tree was built by using PhyML v. 3.0 (Guindon et al. 2010a, b). Bootstrap support values (≥ 70 %) are given above the branches. Mucor simplex CBS 428.80 was used as outgroup. The new species proposed in the present study is indicated in bold. T = ex-type. Eduardo Álvarez Duarte & Constanza Muñoz Cortés, Mycology Unit, Biomedical Sciences Department, University of Chile, Santiago, Chile; e-mail: ealvarezd@med.uchile.cl, camunoz19@outlook.es Ignacio Fernández, Myotis-Chile, Duble Almeyda 2010, Ñuñoa, Santiago, Chile; e-mail: i.fernandezlatapiat@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 336 Persoonia – Volume 45, 2020 Colletotrichum cycadis 337 Fungal Planet description sheets Fungal Planet 1147 – 19 December 2020 Colletotrichum cycadis Andjic, Maxwell & Smith, sp. nov. Etymology. Named after the host genus, Cycas, from which it was isolated. Classification — Glomerellaceae, Glomerellales, Sordariomycetes. Sexual morph not observed. Asexual morph on malt extract agar (MEA) (microscopic preparations in lacto-glycerol, with at least 30 measurements per structure). Hyphae hyaline to pale brown, smooth-walled, septate, branched. Mycelium white, becoming olive grey with age. Conidiomata acervular, brown to black. Conidiophores reduced to conidiogenous cells. Conidiogenous cells hyaline to pale brown, smooth-walled, aseptate, occasionally septate, mostly cylindrical, gradually thinner towards the apex, (7.5 –)11–12(–17.5) × (2 –)2.5 – 3(– 4.5) µm (av. ± SD = 11.5 ± 2.3 × 2.6 ± 0.6 μm, L/W ratio = 4.4). Conidia aseptate, hyaline, smooth-walled, cylindrical to fusoid with obtuse ends, sometimes tapering towards the apex, contents granular or guttulate, 9.5 –13.5 × 3 – 4 μm (av. ± SD = 11.5 ± 0.65 × 3.5 ± 0.2 μm, L/W ratio = 3.3). Appressoria single or in small groups, pale to dark brown, smooth-walled, variable in shape, ovate to irregularly lobed, often tapering towards apex, 4.5 – 8 × 2 – 5.5 μm (av. ± SD = 6.1 ± 1.1 × 4.5 ± 0.9 μm, L/W ratio = 1.3). Culture characteristics — Colonies grown from single conidium on MEA reaching 70–80 mm diam after 10 d at 25 °C in the dark, light grey (5Y 7/1) to light olive grey (5Y 6/2) (Munsell & Munsell 2000); with orange conidial ooze on the surface of colony, aerial mycelium white, tufted near centre. Reverse dark olive grey (5Y 3/2), surrounded with white cottony mycelium. Typus. chiNA, Fujian, Zhangzhou, on leaves of Cycas revoluta, intercepted at Australian border, July 2019, V. Andjic & A. Maxwell (holotype BRIP 71326a, includes holotype culture, LSU, ITS, chs-1, gapdh and tub2 sequences GenBank MW136942, MT439915, MT439917, MT439919, and MT439921, MycoBank MB836054). Additional material examined. chiNA, Fujian, Zhangzhou, on leaves of Cycas revoluta, intercepted at Australian border, July 2019, V. Andjic & A. Maxwell, AQISWA201901 (culture dead), LSU, ITS, chs-1, gapdh and tub2 sequences GenBank MW136943, MT439916, MT439918, MT439920, and MT439922. Notes — Leaf spots were observed on the young leaves of Cycas revoluta in a post entry quarantine greenhouse in Carabooda, Western Australia, Australia. Infected plants were destroyed and the pathogen remains absent from Australia (Australian Plant Pest Database 2020). The leaf symptoms were characterised by chlorosis starting from the tip of the leaf going towards the base where it becomes cream and then dark brown. Conidiomata occur in small, black, irregular shaped aggregates, sometimes clustered in concentric circles. A phylogenetic tree obtained using Bayesian analysis as implemented in Geneious R10 (https://www.geneious.com) of separate and combined sequence data from four gene loci (chs-1, gapdh, ITS and tub2) placed the pathogenic fungus in the Kahawae clade in the C. gloeosporioides species complex (Weir et al. 2012). It is phylogenetically distinct from all other species of the Kahawae clade and can be distinguished with all loci studied, except LSU and tub2. Based on megablast searches on NCBIs GenBank nucleotide database, the closest match to C. cycadis using the LSU sequences was C. lentis (99.83 %), using ITS was C. cobbttiense (98.92 % identity), using chs-1 was C. wuxiense and C. aotearoa (98.62 % identity), and using gapdh was C. aotearoa (98.91 % identity). Colour illustrations. Cycas revoluta plant. Symptomatic leaves; appressoria; conidiogenous cells; colony on MEA at 10 d; conidia. Scale bars = 10 μm. Phylogenetic tree from Bayesian analysis based on combined gene sequences (chs-1, gapdh, ITS and tub2) showing the phylogenetic relationships amongst the newly described taxon C. cycadis (in bold) and known species in the C. gloeosporoides complex. Bayesian posterior probabilities (PP > 0.95) are shown at the nodes. The tree is rooted with C. boninense (ICMP 17904). Ex-type cultures are marked with an asterisk (*). The alignment and tree were deposited in TreeBASE (Submission ID S26714). Vera Andjic, Aaron Maxwell & Keelin Smith, Department of Agriculture, Water and Environment, 24 Fricker Rd., Perth, 6105 Western Australia, Australia; e-mail: vera.andjic@agriculture.gov.au, aaron.maxwell@agriculture.gov.au & keelin.smith@agriculture.gov.au © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 338 Persoonia – Volume 45, 2020 Coprinopsis rubra 339 Fungal Planet description sheets Fungal Planet 1148 – 19 December 2020 Coprinopsis rubra Örstadius, E. Larss. & L. Nagy, sp. nov. Typus. sweDeN, Halland, Steninge, Lövängen, about 15 km N of Halmstad, on cow dung, 28 Aug. 2019, B. Larsson (EL387-19, holotype GB-0207585, ITS-LSU sequence GenBank MT800814, MycoBank MB836567). Etymology. The epithet refers to the red cap and veil colour. Classification — Psathyrellaceae, Agaricales, Agaricomycetes. Additional materials examined. sweDeN, Halland, Varberg, Vadkärr, on cow dung, 17 Aug. 2018, K. Persson (LÖ73-18, GB); Halland, Steninge, Lövängen, about 15 km N of Halmstad, on cow dung, 30 Aug. 2019, B. Larsson & K. Persson (LÖ47-19, GB-0207595); Halland, Mannarp, on cow dung, 18 Sept. 2009, L. Nagy, M. Jeppson & T. Knutsson (NL-2758). Basidiomata small, coprinoid. Pileus at first ellipsoid, campanulate, then expanded convex to plane, umbonate, 6–12 mm wide, radially grooved, red to pale red below the veil, when mature or old pallescent becoming grey tinged; veil vividly red to dark red, covering greater part of surface, splitting into flocci especially at centre. Lamellae free, medium spaced, L = c. 25, when young whitish, becoming brown to blackish, with pale red edge, partly deliquescent. Stipe 12–20 × 1–2 mm, thickened towards base, not root-like extended, concolorous with cap at base, with pale red to whitish upper part, fibrillose, with flocculose veil remnants particularly towards base. Smell not distinctive; taste not recorded. Basidiospores 8 –10 × 4.8 – 5.4 µm (av. 8.8 – 9.2 × 5.1 µm, Qav = 1.7–1.8), oblong, ellipsoid, ovoid, sometimes slightly irregular, in profile flattened on one side, neither amygdaliform nor phaseoliform, rarely broken, in water red (Mu. 2.5YR 4/8, Munsell 1975), with small, central, rather distinct germ pore. Basidia 4-spored, 15 – 30 × 7– 8 µm, surrounded by (3 –)4(– 5) pseudoparaphyses. Pleurocystidia 35 – 65 × 20 – 32 µm, subutriform, ventricose, clavate, sphaeropedunculate, numerous, pale. Cheilocystidia 15 – 50 × 12 – 30 µm, similar to pleurocystidia in shape, ellipsoid, numerous. Pileipellis a cutis made up of hyphae with short, 7–14 µm wide cells. Veil cells 20–80 × 5–30 µm, pale to moderately red intracellular pigmented; surface with dark red spots, irregularly and loosely attached, disappearing when gently tapping on the coverslip. Clamp connections seen at stem base mycelium and veil hyphae. Habitat & Distribution — Growing scattered on cow dung in pastures, only manured from the grazing animals. So far known from three localities in Halland, a southern province of Sweden. Notes — Coprinopsis rubra can be recognised by the striking dark red colour of its cap and veil, coprophilous habitat, and rather small spores. The species belongs to subsection Lanatuli (Uljé 2005) characterised by a hairy-floccose veil made up of elongate elements. Subsection Alachuani differs in having diverticulate often thick-walled elements (Uljé 2005). Coprinopsis erythrocephala is morphologically closely related but can be separated by larger basidiomata, a soon disappearing veil, larger pleurocystidia, larger spores, and a non-coprophilous habitat. In the phylogenetic analysis C. rubra comes out closest to an ITS sequence of an unknown species of Coprinopsis from Brazil, and in the sister clade to C. erythrocephala. Coprinopsis uliginicola KC992960 TYPE 100 Coprinopsis cineraria CKC992962 TYPE Coprinopsis atramentaria AF345814 Coprinopsis jonesii JX118726 82 66 Coprinopsis lagopus FM163179/FM160730 Coprinopsis sp. MK843965 62 77 Coprinopsis rubra MT800814 HOLOTYPE Coprinopsis erythrocephala FN396125/FN396174 84 62 Coprinopsis krieglsteineri FM878019/JX118786 Coprinopsis pannucioides DQ389727 Coprinopsis canoceps KC992964 99 Coprinopsis submicrospora KC992959 TYPE 50 84 Coprinopsis marcescibilis DQ389728 Coprinopsis musae KC992965 TYPE 64 100 Coprinopsis nivea MT889691 Coprinopsis pseudonivea FM163181/FM160728 100 Parasola schroeterii JN943136/HQ847114 Parasola plicatilis FM16321/FM160694 Agrocybe pusiola DQ389732 20.0 Colour illustrations. Sweden, Halland, Steninge, Lövängen, a pasture from the type locality. Basidioma (Varberg, Vadkärr, LÖ73-18); basidiomata (Steninge, holotype); spores; pleurocystidia (above) and cheilocystidia (below); veil cells. Scale bars = 1 cm (basidiomata), 10 µm (spores, cystidia and veil). Phylogram obtained using PAUP* v. 4.0a (Swofford 2003) based on ITS and LSU sequence data showing the position of C. rubra in the Atramentarii and Lanatuli clades (Nagy et al. 2013). Bootstrap values are indicated on branches and the holotype is marked in bold. Leif Örstadius, Lyckans väg 39A, S-29143 Kristianstad, Sweden; e-mail: leif.orstadius@gmail.com Ellen Larsson, Biological and Environmental Sciences, University of Gothenburg, Box 461, 40530 Göteborg, Sweden, and Gothenburg Global Biodiversity Centre, Box 461, SE40530 Göteborg, Sweden; e-mail: ellen.larsson@bioenv.gu.se László G. Nagy, Institute of Biochemistry, Biological Research Center, Temesvari krt 62, H-6726 Szeged, Hungary; e-mail: lnagy@fungenomelab.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 340 Persoonia – Volume 45, 2020 Crinipellis nigrolamellata 341 Fungal Planet description sheets Fungal Planet 1149 – 19 December 2020 Crinipellis nigrolamellata Antonín, Fiard, Ševčíková, Dumez & Courtec., sp. nov. Etymology. The epithet refers to the lamellae that become black with age. Classification — Marasmiaceae, Agaricales, Agaricomycetes. Pileus 5 ‒10 mm broad, convex-conical to broadly conical with a shallow central umbilicus and involute to inflated margin, then broadly conical with central umbo with distinct umbilicus, distinctly radially fibrillose, fibrils projecting up to 1 mm beyond pileus margin, margin shallowly sulcate, dark brown to black-brown (8F3 ‒ 5; Kornerup & Wanscher 1978) at centre, other parts brown-argillaceous or brown (6E5), outermost part paler (± 8D5) or even dirty whitish in old basidiomata. Lamellae moderately close, L = 24‒28, l = 3, emarginate and with small tooth, slightly ventricose, (greenish) grey (± 5D3), with finely pubescent, at first whitish edge; edge and adjacent parts becoming stainy to entirely black. Stipe 15 ‒ 45 × 0.5 ‒1(‒1.5) mm, cylindrical, very slightly broadened at base, insititious, entirely tomentose to adpressedly hairy, strigose at base, sometimes longitudinally striate, entirely dark brown to black-brown (concolorous with pileus centre). Rhizomorphs absent. Basidiospores (7.5 ‒)8 ‒ 9.5(‒10) × (2.7‒)3 ‒ 4 μm, av. 8.7 × 3.5 μm, E = (2 ‒)2.3 ‒ 2.8(‒ 2.9), Q = 2.4 ‒ 2.5, fusoid, lacrimoid, thin-walled, colourless or greyish brownish in KOH, sometimes with one septum, non-dextrinoid. Basidia 17‒ 20 × 6 ‒ 8 μm, 4-spored, clavate; rare sclerobasidia present, black in KOH. Basidioles 13 ‒ 25 × 3 ‒ 9 μm, clavate, subcylindrical, subfusoid. Cheilocystidia 11‒ 27 × 6 ‒ 8.5(‒10) μm, clavate, fusoid, rarely subutriform, mostly with apical projections or (rarely) subcoralloid, less frequently simple, thin- to slightly thick-walled, colourless to often with dark (greyish) blackish contents in KOH. Pleurocystidia (17‒)20‒35 × 6‒9 μm, fusoid, sometimes subrostrate, thin-walled, mostly colourless or pale greyish in KOH. Pileipellis (hypotrichium) a cutis composed of cylindrical, thin- to slightly thick-walled, non-dextrinoid, 3 ‒ 8 μm wide hyphae; pileus hairs up to c. 1300 × 2 ‒ 6(‒ 9) μm, cylindrical, obtuse to subacute, thick-walled (walls up to 1.5(‒ 4) μm), often curved especially at base, often septate or with obliterated lumen, dextrinoid, walls reddish brown in H2O, brown-olivaceous to pale olivaceous in KOH, covered with granular or irregular brown incrustation, more frequently towards base. Stipitipellis a cutis composed of cylindrical, slightly thick-walled, 2 ‒ 5 μm wide, walls ± colourless or pale brownish in H2O; stipe hairs similar to pileus ones, 15 ‒ 600 × 4 ‒17 μm. Clamp connections present. Habit, Habitat & Distribution — Solitary or in groups on fallen leaves in forests. So far known only from Martinique, France. Typus. frANce, Martinique, Trinité com., Tartane, Point Rouge Reserve, Pointe à Bibi, on leaves of Pisonia fragrans (Nyctaginaceae), 3 Nov. 2015, R. Courtecuisse (holotype LIP 0201684, LSU and ITS sequences GenBank MT946361 and MT946363, MycoBank MB836917). Colour illustrations. Locality (France, Martinique, Caravelle NR). From top to bottom: basidiomata; stipe hair. Drawing: basidia, basidiospores, cheilocystidia, pileus hairs, stipe hair, pleurocystidia. Scale bars = 1 cm (basidiomata), 10 μm (all microcharacters). Additional materials examined. frANce, Martinique, Trinité com., Tartane, Caravelle Nature Reserve, Anse Four à Chaux, on fallen leaves, 4 Nov. 2015, V. Antonín & R. Courtecuisse (LIP 0201685, LSU and ITS sequences GenBank MT946362 and MT946364); ibid., on fallen leaves, 17 Dec. 2001, J.P. Fiard F2480 (LIP 0701686). Notes — Crinipellis nigrolamellata is characterised by a dark brown to black-brown pileus and stipe, lamellae becoming black, non-dextrinoid, narrow basidiospores, small cheilocystidia mostly with apical projections, well-developed pleurocystidia, and hairs walls reddish brown in H2O, brown-olivaceous to pale olivaceous in KOH, and covered with brown incrustation. Black coloured lamellae are described in C. bisulcata known from Ecuador and Venezuela. It differs by shorter and differently shaped basidiospores, 6.3 ‒ 8.5 × 3.1‒ 3.8 µm (mostly 7‒ 8.5 × 3‒3.8 µm) and longer, 37‒56 × 4‒7.5 µm, cheilocystidia (Singer 1942). However, Singer mentioned that the black lamellae colour of the type specimen may be caused by a bad preservation – specimens were preserved in alcohol at first and then dried. Crinipellis brunnescens also has lamellae brown to black with age or when where bruised. It differs by a smaller stipe, 8 –12 × 0.4– 0.8 mm, larger basidiospores, 6 –10 × 4 – 5 µm and the absence of pleurocystidia (Kerekes & Desjardin 2009). Other phylogenetically relatively close species never have dark coloured lamellae. Moreover, C. malesiana has a brown to brownish orange pileus at the margin with age, larger basidiospores, longer pleurocystidia, larger, mostly simple cheilocystidia (Kerekes & Desjardin 2009); C. actinophora also differs by a shorter stipe, the presence of rhizomorphs and the absence of pleurocystidia (Singer 1955, Kerekes & Desjardin 2009); C. pallidipilus has golden brown, then pallescent pileus hairs, a shorter stipe, abundant rhizomorphs, larger basidiospores, cheilocystidia with numerous digitate projections and lacks pleurocystidia (Antonín et al. 2014); C. wandoensis differs by well-developed rhizomorphs, broader basidiospores and absent pleurocystidia (Antonín et al. 2014). Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Crinipellis sp. (strain GL-2017, GenBank LT716050.1; Identities = 615/648 (95 %), eight gaps (1 %)) and the type of Crinipellis pallidipilus (strain BRNM 751595, GenBank KF380833.1; Identities = 572/603 (95 %), 11 gaps (1 %)). Closest hits using the LSU sequence are Crinipellis setipes (strain Bandala 4085, GenBank MN567618.1; Identities = 998/ 1021(98 %), 4 gaps (0 %)) and Crinipellis nigricaulis (strain G1325, GenBank MK277894.1; Identities = 997/1021(98 %), 4 gaps (0 %)). Supplementary material FP1149 Phylogram: Best tree from the ML analysis of the nrITS dataset for Crinipellis nigrolamellata and related species with Marasmius crinis-equi as outgroup. Phylogenetic analyses were carried out online at http://phylogeny. lirmm.fr/ (Dereeper et al. 2008) with PhyML v. 3.0 (Guindon et al. 2010a). Multiple sequence alignments were carried out with MUSCLE v. 3.7 (Edgar 2004). Trees were constructed using TreeDyn v. 198.3 (Chevenet et al. 2006) and edited with the newly generated sequences in bold. Vladimír Antonín & Hana Ševčíková, Dept. of Botany, Moravian Museum, Zelný trh 6, 659 37 Brno, Czech Republic; e-mail: vantonin@mzm.cz & hsevcikova@mzm.cz Régis Courtecuisse & Sylvain Dumez, ULR 4515 - LGCgE (Laboratoire de Génie Civil et géo-Environnement), ER4 (Fonctionnement des écosystèmes terrestres anthropisés) - LSVF (Laboratoire des sciences végétales et fongiques), Faculté des sciences pharmaceutiques, Université de Lille, 3, rue du Professeur Laguesse, F-59006 Lille Cedex; e-mail: regis.courtecuisse@univ-lille.fr & sylvain.dumez@univ-lille.fr Jean-Pierre Fiard, 3/524, résidence les Cyclades, Rue R. Garcin, F-97200 Fort-de-France; e-mail: jpfiard@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 342 Persoonia – Volume 45, 2020 Cyberlindnera dauci 343 Fungal Planet description sheets Fungal Planet 1150 – 19 December 2020 Cyberlindnera dauci A.M. Glushakova, M.A. Tomashevskaya & Kachalkin, sp. nov. Etymology. Name refers to Daucus carota from which the species was isolated. Classification — Wickerhamomycetaceae, Saccharomycetales, Saccharomycetes. On glucose peptone yeast extract agar (GPYA) and 5 % malt extract agar (MEA), after 7 d at 25 °C, streak is white-cream, butyrous, with a smooth surface and entire margin. Cells are subglobose, ovoid to elongate (2–5.5 × 2.5–6.5 μm) and occur singly or in pairs, dividing by multilateral budding. After growth on potato dextrose agar (PDA) cells have visible lipid-like body. Ascospores, pseudohyphae and true hyphae have not been observed during 4 wk at 10 and 25 °C in culture (pure cultures and in mating test) grown on GPYA, MEA, PDA, cornmeal agar (CMA), McClary acetate agar and yeast nitrogen base with 0.5 % glucose (YNB) agar. Fermentation of glucose, sucrose and raffinose are positive. Glucose, inulin, sucrose, raffinose, trehalose (delayed weak), cellobiose, salicin, L-rhamnose, D-xylose, ethanol, glycerol, D-mannitol (weak), D-glucitol, DL-lactic acid (weak), succinic acid, citric acid and arbutin are assimilated; no growth occurs on melibiose, galactose, lactose, maltose, melezitose, methyl alpha-D-glucoside, soluble starch, L-sorbose, L-arabinose, D-arabinose, D-ribose, methanol, erythritol, ribitol, galactitol, myo-inositol, D-glucosamine, N-acetyl-D-glucosamine, hexadecane, 2-keto-D-gluconate, 5-keto-D-gluconate and D-glucuronate. Assimilation of nitrogen compounds: positive for ammonium sulfate, cadaverine, creatinine, creatine, L-lysine, D-glucosamine, and negative for potassium nitrate. Growth on vitamin-free medium and on 50 % w/w glucose / yeast extract (0.5 %) agar is positive. Growth on MEA with 10 % NaCl is delayed weak. Growth with 0.01 % and 0.1 % cycloheximide is negative. Starch-like compounds are not produced. Diazonium blue B colour and urease reactions are negative. Maximum growth temperature is 27.5 °C. Typus. russiA, Moscow region, from carrot sample bought on local market, Feb. 2020, A.M. Glushakova, fvmr-2 (holotype KBP Y-6686, preserved in a metabolically inactive state, ex-type cultures VKM Y-3058 = DSM 111207 = CBS 16524, SSU, ITS-D1/D2 domains of LSU nrDNA, tef1 and rpb1 sequences GenBank MT636884, MT636878, LR814018 and LR814019, MycoBank MB836776). Additional material examined. russiA, Moscow region, from carrot sample bought on local market, Feb. 2020, A.M. Glushakova, KBP YE-0693, ITS-D1/ D2 domains of LSU nrDNA sequences GenBank MT936327 and MT939260. Notes — Analysis of the ITS-D1/D2 regions of the surveyed asexual yeasts suggested that they were conspecific and represented a hitherto undescribed species of Cyberlindnera. Based on the NCBI GenBank nucleotide database, the best hit Colour illustrations. Russia, Moscow region, carrots on local market (photo provided by Yu.A. Kachalkina). Cyberlindnera dauci KBP Y-6686: growth of yeast colonies on MEA, yeast cells on PDA and MEA (after 7 d at 25 °C). Scale bar = 5 μm. using the ITS sequence is Cyberlindnera galapagoensis CBS 13997T (GenBank NR_159816; 86.56 % similar, 48 subst. and 29 gaps), using LSU it is Cyb. galapagoensis CBS 13997T (GenBank KJ020281; 96.98 % similar, 17 subst.), using SSU it is Candida mengyuniae CBS 10845T (GenBank EU043157; 96.28 % similar, 48 subst. and 14 gaps), using tef1 it is Cyb. mrakii CBS 1707T (GenBank EU307984; 92.92 % similar, 26 subst. and 4 gaps) and using rpb1 it is Cyb. fabianii YJS4271 (GenBank LK052886; 82.51 % similar, 109 subst. and 1 gap). In compliance with a recent phylogenetic analysis of the Cyberlindnera clade (Zheng et al. 2017), the placement of the new species is demonstrated using the combined ITS and LSU rDNA phylogeny. Cyberlindnera dauci can be physiologically differentiated from the phylogenetically most close species Cyb. galapagoensis based on its ability to assimilate trehalose, cellobiose, L-rhamnose and DL-lactic acid. C. easanensis (HM461688/AY634571) C. maesa (HM461661/JQ812697) Cyb. xylosilytica (KP232976/EF550324) Cyb. japonica (KY103061/EF550323) Cyb. veronae (CBS data/EF550322) 92 Cyb. amylophila (KY103039/EF550319) 100 Cyb. mississippiensis (KY103068/EF550320) Cyb. fabianii (CBS data/EF550321) 77 Cyb. wuzhiensis (FJ606824/FJ606825) 100 Cyb. euphorbiae (KY103041/EF550326) Cyb. meyerae (KY103066/EF550327) C. maritima (KY102197/EF550332) 51 C. mycetangii (KY102221/EF550330) 95 Cyb. americana (KY103037/EF550328) C. nakhonratchasimensis (KY102223/AY634567) 60 Cyb. bimundalis (KY103040/EF550329) C. pattaniensis (HM461657/AY634568) Cyb. xishuangbannaensis (KY213821/KY213813) 55 100 C. stauntonica (HM461658/JQ812698) 0.2 93 61 C. taoyuanica (FJ873419/JQ812699) C. adriatica (KY101825/HE574661) 60 C. hungchunana (HQ623543/JQ812700) Cyb. euphorbiiphila (CBS data/EF550312) C. mengyuniae (EU043159/EU043158) 63 Cyb. samutprakarnensis (AB695388/AB598079) Cyb. tropicalis (KY010353) 100 Cyb. dauci KBP Y-6686T (MT636878) 100 Cyb. dauci KBP YE-0693 (MT936327) Cyb. galapagoensis (KJ020281) Cyb. culbertsonii (KM384448/KM408121) Cyb. misumaiensis (KY103070/U73581) 100 Cyb. lachancei (KY103063/EF550313) C. takata (JQ906769/JQ906764) C. vartiovaarae (KY102489/EF550315) Cyb. subsufficiens (EU307975/EF550318) 99 Cyb. mrakii (EU307973/EF550317) 100 Cyb. saturnus (EU307970/EF550316) 95 Cyb. sargentensis (EU307980/HM461618) Cyb. suaveolens (EU307977/EU544674) Cyb. rhizosphaerae (CBS data) 75 Cyb. xylebori (KY103116/AB534167) Cyb. petersonii (KY103077/EF550311) Cyb. jadinii (DQ249199/EF550309) Cyb. maclurae (KY103065/EF550310) 92 100 75 Maximum likelihood (ML) tree for the Cyberlindnera clade obtained from the combined analysis of ITS and LSU sequence data. The alignment included 1 194 bp and was performed with MAFFT v. 7 (Katoh et al. 2019). The General Time Reversible model (GTR) with Gamma distribution and invariant sites (G+I) was used as the best nucleotide substitution model. The phylogenetic analysis was conducted in MEGA v. 6 (Tamura et al. 2013). Pichia membranifaciens NRRL Y-2026 (DQ104710/U75725) was used as outgroup (hidden). Anna M. Glushakova, Lomonosov Moscow State University, 119234, Moscow, Leninskie Gory Str. 1/12, Russia, and Mechnikov Research Institute for Vaccines and Sera, 105064, Moscow, Maly Kazenny by-street, 5A, Russia; e-mail: glushakova.anya@yandex.ru Maria A. Tomashevskaya, All-Russian Collection of Microorganisms, G.K. Skryabin Institute of Biochemistry and Physiology of Microorganisms RAS, 142290, Pushchino, pr. Nauki 5, Russia; e-mail: tomkotik@rambler.ru Aleksey V. Kachalkin, Lomonosov Moscow State University, 119234, Moscow, Leninskie Gory Str. 1/12, Russia, and All-Russian Collection of Microorganisms, G.K. Skryabin Institute of Biochemistry and Physiology of Microorganisms RAS, 142290, Pushchino, pr. Nauki 5, Russia; e-mail: kachalkin_a@mail.ru © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 344 Persoonia – Volume 45, 2020 Cyphellophora vietnamensis 345 Fungal Planet description sheets Fungal Planet 1151 – 19 December 2020 Cyphellophora vietnamensis Iturrieta-González, Dania García, Guarro & Gené, sp. nov. Etymology. Name refers to the geographical region where the fungus was collected. Classification — Cyphellophoraceae, Chaetothyriales, Eurotiomycetes. Mycelium consisting of branched, septate, subhyaline to pale olivaceous, smooth-walled hyphae, 1–1.5 µm diam. Conidiophores commonly macronematous, mononematous or in groups of 2–4, growing laterally or terminally on hyphae, erect, more or less penicillately branched, up to 250 µm long, with stipe pale brown to brown, smooth- and thick-walled; branches bearing terminally groups of 2–3 phialides, pale brown, asperulate to verruculose; micronematous conidiophores also present, consisting in phialides growing directly or on short supporting cells from vegetative hyphae. Phialides lageniform, 12 – 20 × 2 – 3.5 µm at the broad part, tapering to a long cylindrical neck with a conspicuous collaret slightly darker than the rest of the phialide, pale olivaceous, smooth-walled. Conidia in long unbranched chains (up to 90 conidia), 0(–1)-septate, ellipsoidal to somewhat fusoid, with truncate ends, obovoid when terminal, pale olivaceous, smooth-walled, 4 –7 × 1– 2 µm. Chlamydospores absent. Sexual morph not observed. Culture characteristics — Colonies on potato dextrose agar (PDA) reaching 18 –19 mm diam after 2 wk at 25 °C, brownish grey to grey (4D2/4B1) (Kornerup & Wanscher 1978), final edge olive (2F8), velvety, radially folded, aerial mycelium scarce, irregular margin; reverse olive (2F8). On potato carrot agar (PCA) reaching 18 – 20 mm after 2 wk at 25 °C, olive grey to olive (3D2/3F8), velvety, flat, aerial mycelium scarce, regular margin; reverse olive (2F8). On oatmeal agar (OA) reaching 18–19 mm diam after 2 wk, pale grey to olive (1B1/2F8), velvety at the centre, flat, aerial mycelium scarce, irregular margin; producing a metallic brightness on the border of the colony; reverse olive (2F8). Urease positive; laccase production negative. Cardinal temperatures for growth — Minimum 15 °C, optimum 25 °C, maximum 30 °C. Typus. VietNAm, Northeast region, on unidentified dead leaf, Aug. 2011, J. Guarro (holotype CBS H-24475, cultures ex-type FMR 17714 = CBS 146924; ITS, LSU and tub2 sequences GenBank LR814107, LR814108 and LR814116, MycoBank MB836045). Notes — Based on a megablast search of NCBIs GenBank database, the LSU sequence of C. vietnamensis showed a similarity of 98.22 % (829/844) with the sequence of C. oxyspora (CBS 698.73, GenBank NG_067405) and 97.75 % (825/ 844) with that of C. suttonii (CBS 125441, GenBank MH874978); the ITS sequence was 96.71 % (558/577) similar with that of Phialophora capiguarae (ex-type strain CBS 132767, GenBank KF928464) and a 88.61 % (537/606) with respect to C. oxyspora (IFM 51368, GenBank AB190870); and the tub2 sequence was 94.65 % (336/355) similar with that of P. capiguarae (strain CBS 131954, GenBank KF928593) and a 77.74 % (255/328) with respect to C. ludoviensis (CMRP 1317, GenBank KX583749). Phylogenetic reconstruction with ITS, LSU and tub2 loci (Attili-Angelis et al. 2014) of the accepted species of Cyphellophora and Phialophora, including the type Colour illustrations. Vietnam, Northeast region. Colony sporulating on OA after 2 wk at 25 ºC; conidiophores, phialides and conidia after 18 d. Scale bars = 10 mm (colony), = 10 µm (microscopic structures). species of the respective genera (i.e., C. laciniata CBS 190.61 and P. verrucosa CBS 140326), showed that the new species is allocated in a strongly supported clade with C. oxyspora and P. capiguarae, but being closely related to the latter species. Our phylogeny supports that P. capiguarae as well as P. attinorum, both described by Attili-Angelis et al. (2014), belong to the Cyphellophora clade. Although P. capiguarae was previously considered a species of Cyphellophora (Gomes et al. 2016), the formal taxonomic change was not proposed. Therefore, respective new combinations are proposed below. Morphologically, C. vietnamensis differs from P. capiguarae mainly by having unbranched conidial chains, which are smaller (4 –7 × 1– 2 µm vs 6.5 – 9 × 1.9 – 2.5 µm in P. capiguarae) and commonly aseptate, absence of chlamydospores, and a moderately faster growth (PDA, 18 –19 mm vs 13 –14 mm in P. capiguarae; OA, 18 –19 mm vs 14 –15 mm in P. capiguarae) after 2 wk at 25 °C. Cyphellophora vietnamensis clearly differs from C. oxyspora (Gams & Holubová-Jechová 1976, Réblová et al. 2013) by its long penicillate conidiophores. Cyphellophora attinorum (Attili-Angelis et al.) IturrietaGonzález, Gené, Dania García, comb. nov. — MycoBank MB836046 Basionym. Phialophora attinorum Attili-Angelis et al., ‘attae’ Fungal Diversity 65: 68. 2014. Typus. brAzil, Fazenda Santana, Botucatu, São Paulo, from the cuticle of Atta capiguara gynes, Nov. 2008, A.P.M. Duarte, F.L.A. Guedes & D. AttiliAngelis (holotype and cultures ex-type CBS 131958; ITS, LSU and tub2 sequences GenBank KF928463, KF928527 and KF928591). Notes — Cyphellophora attinorum is closely related to C. livistonae (Crous et al. 2012, Madrid et al. 2016) and C. sessilis (De Hoog et al. 1999, Réblová et al. 2013), both species formerly classified in Phialophora. Morphologically, C. attinorum can be differentiated from C. livistonae by the production of shorter (1.6 – 4.2 vs (4 –)7– 8(–10) µm) and aseptate conidia, and by the absence of chlamydospores. Chlamydospores in C. livistonae are intercalary, 0 –1-septate, measuring 8 –10 × 3 – 5 µm (Crous et al. 2012). Cyphellophora sessilis differs by its shorter (up 3 µm; up to 4.2 in C. attinorum) and obovoidal conidia (broadly ellipsoidal in C. attinorum). Cyphellophora capiguarae (Attili-Angelis et al.) IturrietaGonzález, Gené, Dania García, comb. nov. — MycoBank MB836047 Basionym. Phialophora capiguarae Attili-Angelis et al., Fungal Diversity 65: 70. 2014. Typus. brAzil, Fazenda Santana, Botucatu, São Paulo, from cuticle of Atta capiguara gynes, Dec. 2009, F.C. Pagnocca, N.S. Nagamoto, A.P.M. Duarte & D. Attili-Angelis (holotype and cultures ex-type CBS 132767; ITS, LSU and tub2 sequences GenBank KF928464, KF928528 and KF928592). Supplementary material FP1151 Maximum likelihood tree obtained from the combined analysis of ITS, LSU and tub2 sequences of the genus Cyphellophora and representative species of the genus Phialophora. New species and new combinations proposed are indicated in bold face. Isabel Iturrieta-González, Josepa Gené, Josep Guarro & Dania García, Mycology Unit, Medical School and IISPV, Universitat Rovira i Virgili (URV), Sant Llorenç 21, 43201 Reus, Tarragona, Spain; e-mail: isabeliturrieta@gmail.com, josepa.gene@urv.cat, josep.guarro@urv.cat & dania.garcias@urv.cat © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 346 Persoonia – Volume 45, 2020 Elaphomyces nemoreus 347 Fungal Planet description sheets Fungal Planet 1152 – 19 December 2020 Elaphomyces nemoreus Jeppson, Molia & E. Larss., sp. nov. Etymology. Name refers to the occurrence in deciduous woodlands. Classification — Elaphomycetaceae, Eurotiales, Eurotiomycetes. Ascomata subglobose 1–6 cm diam. Peridial surface yellowish grey-brown with a covering of low and flat, obtuse, somewhat darker warts or platelets on a lighter background. Ascomata are found solitary or in small groups and are covered by a pale yellow to sulphur yellow mycelial layer encrusting soil. Cortex and mycelial covering constructed of loosely to intricately interwoven thin-walled, hyaline to yellowish hyphae, 2 – 5 µm diam, sometimes with slightly encrusted walls. Areas between the warts are formed by compacted parallel bundles of hyaline compacted hyphae. Peridium in section thick (fresh up to 5 mm), distinctly marbled with large ochraceous to dark brown or purplish brown and irregularly rounded marbles divided by winding, more or less radially arranged whitish veins. Peridium formed by loosely to intricately interwoven hyphae up to 10 µm diam, in darker areas of the marbling with adhering brown grains of extra-cellular pigments. Gleba young greyish white, web-like, later pulverulent and black. Asci not observed. Ascospores dark brown, globose, in KOH 3 % 23 – 32 µm (av. 26 µm) including ornamentation, 17– 25 µm (av. 22 µm) ornamentation excluded, in Hoyer´s solution significantly smaller: 20–25.5 µm (av. 22 µm) including ornamentation, 16.5–22 µm (av. 18.7 µm) ornamentation excluded. Ornamentation in side view with broad spines and warts up to 3 µm high. A surface view reveals groups of spines with confluent apices, with age coalescing to form coarse meshes and crests. Habitat & Distribution — Found associated with Fagus sylvatica and Quercus robur on basic and acidic soils. Likely to have a northern distribution range in Europe, but occurs also in Southern Europe. Typus. sweDeN, Bohuslän, Valla, Sundsby, deciduous woodland under Quercus robur and Fagus sylvatica, 20 m asl., N58.065348° E11.676246°, 17 July 2020, E. Larsson & M. Jeppson 11077 (holotype GB-0207587, ITSLSU sequence GenBank MT872017, MycoBank MB837347). Additional materials examined. Elaphomyces decipiens: sweDeN, Gotland, Mästerby, wooded meadow under Quercus robur, 24 Oct. 2019, E. Larsson 268-19 (GB-0207592), ITS-LSU sequence GenBank MT872011. Elaphomyces nemoreus: NorwAy, Agder, Farsund, 2013, A. Molia 351-2013 (O-F21484), ITS-LSU sequence GenBank KR029742; Aust-Agder, Arendal, 9 Nov. 2013, A. Molia et al. (O-F21513), ITS-LSU sequence GenBank KR027943. – sweDeN, Bohuslän, Valla, Sundsby, deciduous woodland under Fagus sylvatica, 26 Aug. 2014, K. Rense & M. Jeppson 10151 (GB-0207593), ITS-LSU sequence GenBank MF614923; ibid., 17 July 2020, E. Larsson & M. Jeppson 11180 (GB); Bohuslän, Ljung, Tjöstelseröd, under Quercus robur, 24 Apr. 2020, E. Larsson & M. Jeppson 11141 (GB), ITS sequence GenBank MT872012; ibid., 24 Apr. 2020, E. Larsson & M. Jeppson 11139 (GB-0207589); Bohuslän, Resteröd, under Fagus sylvatica, 20 Nov. 2019, E. Larsson 382-19 (GB-0207590), ITS-LSU sequence GenBank MT872015; ibid., 16 Apr. 2020, E. Larsson 44-20 (GB-0207591), ITS-LSU sequence GenBank MT872013; Bohuslän, Uddevalla, Rimnersvallen, deciduous woodland under Quercus robur, 3 Aug. 2016, A. Molia & M. Jeppson 10482 (GB0207594), ITS sequence GenBank MT872016; Västergötland, V. Tunhem, deciduous woodland under Quercus robur, 26 Dec. 2019, A. Bohlin, E. Larsson & M. Jeppson 11076 (GB-0207588), ITS sequence GenBank MT872014. Notes — Elaphomyces nemoreus belongs to Elaphomyces section Elaphomyces subsection Muricati. It is closely related to E. decipiens (with a neotype recently designated by Paz et al. 2017), with which it shares the characteristic marbled peridium with whitish, more or less radially arranged veins and the ochraceous to dark purplish brown, irregularly rounded, large marbles. Both species have a cortex surface with low flat, greybrown warts. In E. nemoreus the surface warts typically appear as plates forming a cheetah pattern. In E. decipiens the hyphal crust surrounding the ascomata is creamy white whereas in E. nemoreus it has distinct sulphur yellow tinges. The spores are similar in size and ornamentation in the two species. In Molia et al. (2020), a genetic divergence was observed within E. decipiens. Further sequenced collections from Scandinavia confirmed this observation and the occurrence of two genetically distinct but morphologically similar species. So, we here recognise E. nemoreus as a distinct species in the subsection Muricati. In the phylogenetic analyses it comes out with support as a sister species to E. decipiens from which it differs by five substitutions and two 1– 2 bp insertion/deletion events in the ITS1 region and two substitutions in the ITS2 region. Based on the sequences included here no gene flow between the two genotypes can be observed, which supports their evolutionary autonomy. The sequences originating from North America submitted to GenBank as E. decipiens (EU837299, EU846311) did not come out together with the neotype of E. decipiens nor with the herein described species, and these two sequences are shown to be more closely related to E. barrioi and the recently described E. bucholtzii (Crous et al. 2020a). Elaphomyces nemoreus is recorded from coastal areas of south-western Scandinavia (Norway and Sweden) where it occurs in south-facing, warm forest habitats with Fagus sylvatica and Quercus robur, often on acid, but more nutrient-rich soils. Elaphomyces nemoreus is more frequently encountered in Scandinavia than E. decipiens, that must be regarded as rare but with confirmed finds from meadow areas under Quercus robur on calcareous ground. Elaphomyces decipiens may have a more southern European distribution range, but sequence data published in Paz et al. (2017) indicate that also E. nemoreus occurs under Fagus in northernmost Spain. Supplementary material Colour illustrations. Elaphomyces nemoreus (holotype), habitat. Ascomata; ascospores in side view and surface view. Scale bars = 10 µm (ascospores), 50 mm (ascomata). FP1152 Phylogram obtained using PAUP v. 4.0a (Swofford 2003) based on ITS and LSU data showing the position of E. nemoreus in Elaphomyces subsection Muricati. Bootstrap values are indicated on branches, E. nemoreus is marked in bold and the holotype is indicated. Mikael Jeppson & Ellen Larsson, Biological and Environmental Sciences, and Gothenburg Global Biodiversity Centre, University of Gothenburg, P.O. Box 461, SE-40530 Göteborg, Sweden; e-mail: mikael.jeppson@bioenv.gu.se & ellen.larsson@bioenv.gu.se Anne Molia, Alette Iversens gate 5, N-3970 Langesund, Norway; e-mail: anne.molia@nhm.uio.no © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 348 Persoonia – Volume 45, 2020 Exophiala embothrii 349 Fungal Planet description sheets Fungal Planet 1153 – 19 December 2020 Exophiala embothrii Sand.-Den. & Giraldo López, sp. nov. Etymology. Named after the plant genus on whose rhizosphere the fungus was isolated, Embothrium. Classification — Herpotrichiellaceae, Chaetothyriales, Chaetothyriomycetidae, Eurotiomycetes. Mycelium consisting of hyaline to pale brown, smooth, branched, septate, 1– 3 µm diam hyphae, torulose hyphae seldom present. Conidiophores short, erect, cylindrical, septate, poorly differentiated 18 – 50 μm long, 1– 2.5 μm wide at the widest portion, often reduced to conidiogenous cells born laterally on the hyphae. Conidiogenous cells terminal or lateral on conidiophores and hyphae, subcylindrical to doliiform, 3.5 –12.5 × 1– 3 μm, or more commonly as lateral pegs borne terminal or intercalary on undifferentiated hyphae, 0.5 –1.5 × 0.5 –1 μm. Conidia aseptate, (sub)hyaline, ellipsoidal to cylindrical, (2.5–)4.5 – 6(–7) × 1.5 – 2.5 μm, often forming palisades alongside the hyphae or small heads on the tip of conidiophores. Chlamydospores and budding cells not observed. Culture characteristics — Colonies on malt extract agar (MEA) dull green to olivaceous grey, velvety to cottony, slightly raised to umbonate, margin entire. Reverse olivaceous black without diffusible pigment. Typus. chile, Los Lagos Region, Osorno, from rhizosphere of Embothrium coccineum (Proteaceae), 1 Jan. 2019, A. Giraldo & N. Sandoval-Giraldo (holotype CBS H-24520, culture ex-type CBS 146558, ITS, LSU, tef1 and tub2 sequences GenBank MW045817, MW045821, MW055980 and MW055976, MycoBank MB837535). Additional materials examined. chile, Los Lagos Region, Osorno, from rhizosphere of Embothrium coccineum, 1 Jan. 2019, A. Giraldo & N. Sandoval-Giraldo, CBS 146559 ITS, LSU, tef1 and tub2 sequences GenBank MW045818, MW045822, MW055981 and MW055977; ibid., CBS 146560, ITS, LSU, tef1 and tub2 sequences GenBank MW045819, MW045823, MW055982 and MW055978; ibid., CBS 146561, ITS, LSU, tef1 and tub2 sequences GenBank MW045820, MW045824, MW055983 and MW055979. Similarly, two other genetically related Exophiala species of the salmonis-clade, E. radicis and E. tremulae, are known to inhabit plant roots of Microthlaspi perfoliatum (Brassicaceae) and Populus tremula (Salicaceae), respectively (Crous et al. 2011, Maciá-Vicente et al. 2016). A four-gene phylogeny based on ITS, LSU, tef1 and tub2 sequences showed that E. embothrii is phylogenetically closely related to E. opportunistica and E. bonariae. However, E. embothrii differ by its consistently more elongated conidia and phialides, the absence of budding cells and the scarce presence of moniliform hyphae. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Exophiala opportunistica (strain CBS 637.69, GenBank JF747121.1; Identities = 549/549 (100 %), 0 gaps), and Exophiala opportunistica (strain CBS 122269, GenBank JF747124.1; Identities = 549/549 (100 %), 0 gaps). Closest hits using the LSU sequence are Exophiala psychrophila (strain CBS 191.87, GenBank MH873750.1; Identities = 812/815 (99 %), no gaps), Exophiala bonariae (strain CCFEE 5899, GenBank KR781082.1; Identities = 812/815 (99 %), no gaps), and Exophiala cancerae (strain CBS 115142, GenBank MH874540.1; Identities = 814/816 (99%), 1 gap (0%)). Closest hits using the tef1 sequence had highest similarity to Exophiala opportunistica (strain CGMCC:3.17515, GenBank KP347908.1; Identities = 192/201 (96 %), no gaps), Exophiala opportunistica (strain CGMCC:3.17507, GenBank KP347907.1; Identities = 192/201 (96 %), no gaps), and Exophiala cancerae (strain CBS 117491, GenBank JN128799.1; Identities = 173/201 (86 %), 3 gaps (1 %)). Closest hits using the tub2 sequence had highest similarity to Exophiala opportunistica (strain CBS 112269, GenBank JN112487.1; Identities = 408/408 (100 %), no gaps), and Exophiala opportunistica (CBS 637.69, GenBank JN112490.1; Identities = 405/408 (99 %), no gaps). Notes — Exophiala embothrii clusters within the salmonisclade of Exophiala. This clade includes typically waterborne mesophilic species, some of which are associated with superficial and in some cases invasive infections mostly on cold-blooded animals, but also including agents of infection in humans and other homeothermic animals (De Hoog et al. 2011, Najafzadeh et al. 2018, Garzon et al. 2019). The new species described here was isolated from the rhizosphere of a native South American Proteaceae species, Embothrium coccineum. Colour illustrations. Embothrium coccineum (‘Notro’ or ‘Chilean firetree’) with the Osorno volcano on the background (photo by Samuel Troncoso Sandoval, from Wikimedia Commons, license CC BY-SA 3.0). Conidiophores; conidiogenous cells; conidia. Scale bars = 5 µm. Maximum likelihood tree (RAxML, conducted in the CIPRES science gateway) from the analysis of combined ITS, LSU, tef1 and tub2 sequences (total 1 957 bp) of members of the Exophiala salmonis clade. Bootstrap support values above 70 % are indicated on the nodes. The new species proposed in this study is indicated in bold. T and IT denotes ex-type and ex-isotype cultures. GenBank reference sequence accession numbers for ITS, LSU, tef1 and tub2 are indicated between parentheses. The tree is rooted to Cladophialophora modesta (CBS 985.96). Marcelo Sandoval-Denis, Westerdijk Fungal Biodiversity Institute, Uppsalalaan 8, 3584 CT Utrecht, The Netherlands; e-mail: m.sandoval@wi.knaw.nl Alejandra Giraldo, Radboud University Medical Centre, Geert Grooteplein Zuid 10, 6525 GA Nijmegen, The Netherlands; e-mail: Alejandra.GiraldoLopez@radboudumc.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 350 Persoonia – Volume 45, 2020 Geoglossum jirinae 351 Fungal Planet description sheets Fungal Planet 1154 – 19 December 2020 Geoglossum jirinae V. Kučera, Ševčíková, Slovák, sp. nov. Etymology. The name ‘jirinae’ honours the collector of the holotype, Jiřina Hrabáková. Classification — Geoglossaceae, Geoglossales, Geoglossomycetes. Ascomata solitary, scattered, clavate, stipitate, 18–36 × 2–4 mm, dry, black. Ascigerous part lanceolate or broadly clavate, 1/3 – 2/3 of the total ascomata length, black, compressed in cross section, clearly delimited from the stipe, smooth both in fresh and dry conditions. Stipe cylindrical, oval in cross section, 9–15 × 1.5–2.5 mm, robust, with black squamules, slightly thickened upward. Asci cylindrical to clavate, (123 –)135 –145(–166) × 14 –17 μm (all measurements of microscopic characters refer to material examined in 3 % KOH), Q = 7.8–9.7, unitunicate, inoperculate, 8-spored, with euamyloid ascoapical apparatus and inamyloid wall in MLZ and IKI. Ascospores elongated clavate to ellipsoid baculiform, usually slightly curved, (39 –) 45 – 57(– 60) × 5 – 6(– 6.5) μm, Q = 7–10(–12), first hyaline, finally becoming brown in water, blackish in 5 % KOH, 1– 4(–7)-septate when mature, most often with three septa, smooth. Ascoconidia not observed. Paraphyses numerous, longer than asci, straight, sparsely or moderately septate, 2 – 3 μm wide, hyaline, agglutinated by light brown amorphous matter in apical part. Apical cells of paraphyses variable, cylindrical, clavate to capitate, curved, contorted, sometimes bifurcate, or proliferating, mostly 22–47 × 2–3 μm, some cells inflated up to 10 μm. Stipe surface squamulose. Hyphae of the squamules straight, moderately septate, formed by chains of several (4 –7) pale brown cells, apical cells clavate. Habit, Habitat & Distribution — Solitary, on soil among grass. The species is known only from the type locality. Notes — The combination of characters involving short ascospores (45 – 55 × 5 – 6 μm) with predominantly three septa (occasionally 0 –7) and stipe with scales, and long (22 – 45 × 2 – 3.5 μm) slightly curved last cell of paraphyses is unique for this Geoglossum species. Macromorphologically similar G. fallax differs in longer (65 –105 × 5 –7 μm) and more septate (7–12) spores (Durand 1908). The steppe habitat on calcareous soil could host also G. cookeanum which is different in chainforming apical cells of paraphyses, almost smooth stipe and 7-septate ascospores (Minter & Cannon 2015). Very close in having a squamulose stipe, spores (50–60 × 4–6 μm) with 1–3 (5–7 when mature) septa is G. vleugelianum, but the difference is in the coloured and stout, upwardly clavulate paraphyses with pyriform or globose apical cells and easily removed tufts of hyphae on the stipe; G. elongatum has likewise elongate paraphyses and relatively short spores (50 – 60 × 5 –7 μm) with 0 –7 septa (Nannfeldt 1942), but has setose hairs on the stipe and therefore was relocated to Hemileucoglossum (Arauzo & Iglesias 2014). It was impossible to verify the type specimen of G. elongatum due to undergoing renovation of the fungarium building (S), but the presence of setose hairs on the stipe is the basic character of the genus Hemileucoglossum. Possibly similar could also be G. fumosum with a densely squamulose stipe, short spores (30 – 40 × 4.5 – 5.5 μm) and asci (100 –125 × 12 –17 μm), but the ascigerous part characteristically looks like it is impregnated by brown smoke (Hakelier 1967). Typus. czech republic, Hrubšice village, Nad řekami Nature Reserve, N49°05'35" E16°17'33", elev. 257 m, on soil in dry steppe lawns on serpentine slopes, 16 Nov. 2019, J. Hrabáková, (holotype SAV F-11578, ITS and LSU sequences GenBank MT940893 and MT940893, MycoBank MB837371). Colour illustrations. Steppe lawns on serpentine slopes near Hrubšice village in the Czech Republic. Macro- and microscopic structures of holotype: ascomata; ascospores (in KOH); amyloid reaction of the ascoapical apparatus (in IKI); paraphyses (in 3 % KOH); stipe surface (in 3 % KOH). Scale bars = 1 cm (ascomata), 10 μm (microscopic structures). The Bayesian majority-rule consensus tree was inferred from the concatenated dataset of ITS-LSU sequences. The dataset included G. jirinae (H: holotype), relevant Geoglossum species, and Leucoglossum leucosporum as an outgroup (TreeBASE study S26857). Bayesian inference was run in MrBayes v. 3.2.7a, using four independent chains, 10 M generations, and a sampling frequency of 1 000 (Ronquist et al. 2012). The best-fit partitioning schemes and models were estimated for the concatenated tree, using the greedy search mode as implemented in the PartitionFinder v. 2.1.1 (Lanfear et al. 2016). The maximum likelihood analysis was computed in RAxML v. 8.2.12 (Stamatakis 2014). Analyses were computed in the CIPRES Science Gateway v. 3.3 (Miller et al. 2010). Numbers above branches indicate Bayesian posterior probabilities ≥ 0.95 and the maximum likelihood bootstrap support values ≥ 85 %. The scale bar represents the number of nucleotide changes per site. Viktor Kučera, Plant Science and Biodiversity Centre, Institute of Botany, Slovak Academy of Sciences, Dúbravská cesta 9, 845 23, Bratislava, Slovakia; e-mail: viktor.kucera@savba.sk Marek Slovák, Plant Science and Biodiversity Centre, Institute of Botany, Slovak Academy of Sciences, Dúbravská cesta 9, 845 23, Bratislava, Slovakia, and Department of Botany, Charles University, Benátská 2, 128 01 Praha, Czech Republic; e-mail: marek.slovak@savba.sk Hana Ševčíková, Department of Botany, Moravian Museum, Zelný trh 6, 659 37, Brno, Czech Republic; e-mail: hanyzka@mail.muni.cz © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 352 Persoonia – Volume 45, 2020 Gymnascella minnisii 353 Fungal Planet description sheets Fungal Planet 1155 – 19 December 2020 Gymnascella minnisii Adam, Rea-Ireland, Smyth & Overton, sp. nov. Etymology. Named after Andrew Minnis who first mentioned the specimen as 24MN30 (GenBank JX270629) in a survey of Eastern United States bat hibernacula in 2013 after the initial outbreak of White-nose syndrome of bats. Classification — Gymnoascaceae, Onygenales, Eurotiomycetes. On oatmeal salt sediment agar: Ascomata gymnothecial-like, more hyphal than peridial, solitary, globose, measuring 28 – 118.5 (av. = 56.80, n = 40) µm diam; yellow grey (3B6; Kornerup & Wanscher 1978); developing slowly and ripening within 90 d at 25 °C (12 h white fluorescent light / 12 h dark). Ascomatal initials clavate with thin curled hyphae; peridial hyphae light orange to gold yellow (5A4 – 5B7), smooth and septate with distinct appendages measuring 3.8– 31 (av. = 14.86, n = 11) × 1.7–4. (av. = 1.93, n = 11) µm. Asci globose to ovoid, 8-spored, 6.8 –10 (av. = 8.5, n = 11) µm diam. Ascospores globose to hat- or saturn-shaped, measuring 2.6 – 4 (av. = 3.34, n = 19) × 2.8 – 3.5 (av. = 3.13, n = 13) µm. Culture characteristics — On Sabouraud dextrose agar (SAB) acidified with 120 µL 85 % lactic acid for optimal pigment production, (12 h white fluorescent light / 12 h dark at 25 °C): Colony yeast-like, at first yellow-grey (3A3–3B6), in age darkening slightly after 90 d. On synthetic nutrient-poor agar (SNA), colony filamentous, at first white to pale yellow-white on SNA (2A1– 2A2). Typus. USA, Pennsylvania, Blair County, Canoe Creek State Park, Canoe Creek Hartman Mine, from bat guano, 15 Mar. 2012, B. Overton LHU G3 (dried, non-metabolically active holotype CUP-70725, in Cornell University Plant Pathology Fungarium, metabolically active culture CBS 147160, ITS sequence GenBank MT988379, MycoBank MB836835). Notes — Gymnascella minnisii can be differentiated from other species of Gymnascella due to the absence of conidial development. The ascomata of G. minnisii are more hyphal and less developed than the peridial ascomata described in Pseudogymnoascus species. Additionally, the hat- or saturn-shaped ascospores place this species in the Onygenales. Genetic analysis of the ITS gene of G. minnisii suggests that the new species described here is identical to the ITS gene sequence of isolate 24MN30 (Lorch et al. 2013) deposited in GenBank (accession number JX270629). Isolate 24MN30 has remained an undescribed species since the publication of their work. This work is the first to unite 24MN30 and LHU G3 under one name through morphological and molecular data. This species forms racket hyphae similar to that observed by Peck (1985). Gymnascella marginispora (MH860180.1) rDNA 676 characters ML Tree PAUP 4.0a.167 -ln L: 1865.663 ML/MP Bootstraps Gymnascella sp. (AY304511.1) 92/98 Gymnascella hyalinospora (KP231285.1) 95/100 Gymnoascus halophilus (KP278211.1) 88/94 Gymnascella aurantiaca (KC009379.1) - /90 Gymnascella minnisii sp. nov. 100/100 Gymnascella sp. (JX270629.1) Rollandina capitata (LC146760.1) 0.01 substitutions/site Colour illustrations. Background photo of Canoe Creek Hartman Mine. Colony front colour on SAB at 90 d; colony back colour on SAB at 90 d; dissecting scope image of ascomata on SNA; DIC racket hyphae; fluorescence image of claw-like initials in calcofluor white; ascospores displaying hat- or saturn-shape in calcofluor white. Scale bar = 100 μm (ascomata), 2 μm (ascomatal initial), 5 μm (all others) Phylogenetic placement of Gymnascella minnisii compared to close relatives on a maximum likelihood tree with maximum likelihood/maximum parsimony bootstrap support values. Gymnascella minnisii is highlighted in bold. This analysis was based on a single gene alignment, utilising nrDNA sequences (ITS1, ITS4 primers; White et al. 1990) only. PAUP v. 4.0a build 167 (Swofford 2003) was utilised to conduct the 1 000 bootstrap maximum parsimony analysis and maximum likelihood analysis. The maximum likelihood analysis utilised the General Time Reversible (GTR) nucleotide model with rate matrix set to estimate, and variable sites set to gamma distribution. Bootstrap support values greater than 70 % are shown on nodes in the following order: maximum likelihood/maximum parsimony. The alignment was deposited in TreeBASE (submission S26902). Jacob D. Adam, Alden J. Mileto & Barrie E. Overton, 205 East Campus Science Center, Lock Haven University, Lock Haven, PA 17745 USA; e-mail: jacob.d.adam.42@gmail.com, ajm1653@lockhaven.edu & boverton@lockhaven.edu Abigail E. Rea-Ireland, University of Tennessee, Knoxville. Knoxville, TN, 37996 USA; e-mail: abbyliz52@gmail.com Christopher W. Smyth, Binghamton University, Binghamton, NY, 13902 USA; e-mail: chris.smyth.psu@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 354 Persoonia – Volume 45, 2020 Pseudogymnoascus palmeri 355 Fungal Planet description sheets Fungal Planet 1156 – 19 December 2020 Pseudogymnoascus palmeri Rea-Ireland, Smyth, Lindner & Overton, sp. nov. Etymology. Named after Jonathan M. Palmer, formerly of the United States Forest Service, for his many contributions to the study of Pseudogymnoascus and his contributions to establishing mating-type genes for the genus. Classification — Pseudeurotiaceae, Thelebolales, Leotiomycetes. On Rose Bengal Agar (RBA): Ascomata gymnothecial, solitary, globose, measuring 85.3 –172.5 (av. = 125.3, n = 10) × 60 –161.8 (av. = 112.9, n = 10) μm in size; grey orange (5B3-6; Kornerup & Wanscher 1978); developing rapidly and ripening within 10 d at 25 °C (12 h white fluorescent light / 12 h dark). Ascomatal initials coiled to irregular; peridium is a gymnothecium composed of textura intricata, the peridial hyphae darkly pigmented brownish yellow (5C7), smooth to minutely roughened with distinct appendages. Asci ovoid, 8-spored, 6.7– 9.5 (av. = 7.6, n = 35) × 4.9 –7.3 (av. = 5.9, n = 35) μm. Ascospores aseptate, fusoid to ellipsoid, smooth, grey orange (5B36-6); 2.9 – 4.1 (av. = 3.4, n = 80) × 1.8– 3 (av. = 2.3, n = 80) μm. Culture characteristics — (12 h white fluorescent light / 12 h dark at 25 °C): Colony colour analysed on Sabouraud dextrose agar (SAB) acidified with 120 µL 85 % lactic acid for optimal pigment production rather than RBA because the pink colour of the agar compromises interpretation of fungal pigmentation. Colony reverse at first yellow white (4A2), maturing to grey orange (5B3-6) with age after 10 d. On oatmeal salt sediment agar, colony reverse colour is diffuse light orange to orange (6A5-7). Typus. USA, Pennsylvania, Centre County, Woodward Cave, from sediment, 2019, B. Overton LHU 407 (dried, non-metabolically active holotype CUP-70724, in Cornell University Plant Pathology Fungarium, metabolically active culture CBS 147159 in the CBS Collection of the Westerdijk Fungal Biodiversity Institute ITS, rpb2, and tef1 sequences GenBank MT988150, MW054468, MW054467; MycoBank MB837413). Notes — Pseudogymnoascus palmeri produces sexual structures on SNA and RBA in the presence of a bacterium co-isolated from the original sediment sample. A BLAST search of the bacterial co-isolate’s 16S rDNA provided a 100 % match with Pseudomonas moorei. Culture became sterile after removal of the co-isolated bacterium using SAB acidified with 120 µL 85 % lactic acid. Sterility was maintained, even when the fungal isolate was re-plated onto RBA or SNA. Morphological analyses suggest that P. palmeri and P. roseus could be sister taxa. They are similar in the morphological characteristics of gymnothecial ascomata production and ascospore size. Samson (1972) described P. roseus as being characterised by pinkish to reddish ascomata, roughened appendages with spines or warts, and the presence of aleurioconidia. Pseudogymnoascus palmeri can be distinguished from P. roseus based on conidiogenesis (P. palmeri does not produce conidia) and colour (P. palmeri ascomata are grey orange). Samson (1972) did not describe Colour illustrations. Background photo of Woodward Cave, Pennsylvania, USA. Confocal laser-scanning image of gymnothecium; DIC image of ascospores on synthetic nutrient-poor agar (SNA); colony back colour on SAB at 10 d; ascomatal initials on SNA at 10 d; asci and peridial hyphae on SNA. Scale bars = 20 µm (gymnothecium), 2 µm (ascomatal initial), 5 µm (all others). the reverse colour of colony plates, but as a morphological character in Pseudogymnoascus, this should not be ignored (Crous et al. 2019c). Minnis & Lindner (2013) were the first to examine many Pseudogymnoascus taxa using modern phylogenetic methods. This work builds off their multi-gene approach, utilising three of the five phylogenetically informative loci useful for phylogenetic species resolution within the genus Pseudogymnoascus (Minnis & Linder 2013). The three-locus phylogenetic analysis conducted in this study indicates strong support for the placement of P. palmeri (LHU 407) in a clade with isolate WSF 3629 (Minnis & Linder 2013). Phylogenetically, WSF 3629 is closely related to clade G in the P. roseus complex (Palmer et al. 2014). This isolate is also of significant interest due to its phylogenetic proximity to the white-nose syndrome pathogen, P. destructans. The relationship of this complex to P. destructans has not been fully resolved, even with a five-gene analysis (Minnis & Linder 2013). WSF 3629 was suggested as a new species by Palmer et al. (2014) but has remained an undescribed species since the publication of their work. This study honours the work of Palmer, and formally describes the new species as Pseudogymnoascus palmeri sp. nov. and identifies a new strain of this species (LHU 407) from Pennsylvania. In addition to morphology, phylogenetic analysis of a three-gene multi-locus alignment (ITS nrDNA, rpb2 and tef1) support the description of isolates LHU 407 and WSF 3629 as the phylogenetic species P. palmeri, distinct from clade G in the P. roseus complex. This study generated the three-locus dataset for isolate LHU 407. The challenge in resolving the clade G in the P. roseus complex, as well as its association with P. destructans, highlights the need for greater biodiversity sampling of the genus Pseudogymnoascus. The multi-gene data for isolate WSF 3629, and the remainder of the Pseudogymnoascus species, were derived from Minnis & Linder (2013), as well as previous species descriptions for P. lindneri and P. turneri (Crous et al. 2019c). The outgroup taxon was Gymnascella minnisii, GenBank rpb2 and tef1 sequences MW054470, MW054469, also described in this issue of Fungal Planet. Supplementary material FP1156 Phylogenetic placement of Pseudogymnoascus palmeri sp. nov. on a strict consensus maximum parsimony tree with maximum likelihood/ maximum parsimony bootstrap support values (based on 1 000 bootstrap pseudo-replicates), was determined from analysis of a multi-gene alignment of rDNA (primers ITS1, ITS4; White et al. 1990), rpb2 (primers RPB2-7cF, RPB2-11aR; Liu et al. 1999), and tef1 (primers EF1-983F, EF1-2218R; Rehner & Buckley 2005). PAUP v. 4.0a build 167 (Swofford 2003) was used to conduct the maximum parsimony analysis. The parsimony analysis generated a single most parsimonious tree which was also the strict consensus. A maximum likelihood analysis was completed using GARLI v. 2.01 (Zwickl 2006) on the CiPRES Science Gateway (Miller et al. 2010). A consensus tree was generated from a single replicate ML analysis with 1 000 bootstrap pseudo-replications. There were no significant topological differences between the parsimony and likelihood consensus trees. For maximum likelihood, the General Time Reversible (GTR) evolutionary model was utilised, the proportion of invariant sites was set to estimate, and the model of rate heterogeneity was set to gamma distribution. Bootstrap support values located at nodes are: Maximum Likelihood/Maximum Parsimony. Alignment and tree(s) are deposited in TreeBASE (study 27014). Abigail E. Rea-Ireland, University of Tennessee, Knoxville, TN, 37996 USA; e-mail: abbyliz52@gmail.com Christopher W. Smyth, Binghamton University, Binghamton, NY, 13902 USA; e-mail: chris.smyth.psu@gmail.com Daniel L. Lindner, One Gifford Pinchot Drive Madison, WI, 53726 USA; e-mail: daniel.l.lindner@usda.gov Brent J. Sewall, Department of Biology, Temple University, 1900 N. 12th Street, Philadelphia, PA, 19122 USA; e-mail: bjsewall@temple.edu Barrie E. Overton, 205 East Campus Science Center, Lock Haven University, Lock Haven, PA, 17745 USA; e-mail: boverton@lockhaven.edu © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 356 Persoonia – Volume 45, 2020 Hypoxylon hepaticolor 357 Fungal Planet description sheets Fungal Planet 1157 – 19 December 2020 Hypoxylon hepaticolor J. Fourn. & A.N. Mill., sp. nov. Typus. freNch GuiANA, Maripasoula, Saül, shortcut toward the airfield, disturbed secondary rainforest, on a dead corticated branch 1.5–2 cm diam, N3.622159 W53.204166, c. 190 m, 20 June 2019, J. Fournier, GYJF 19127 (holotype LIP, isotypes HAST, ILLS00121426, ITS and LSU sequences GenBank MT799854 and MT799853, MycoBank MB837614). Etymology. From Latin hepar, hepatis = liver and color = colour, for the reddish brown stromatal surface colour reminiscent of that of liver. Classification — Hypoxylaceae, Xylariales, Sordariomycetes. Additional materials examined. freNch GuiANA, Maripasoula, Saül, shortcut toward the airfield, disturbed secondary rainforest, on bark, N3.623524, W53.206542, c. 200 m, 17 June 2019, J. Fournier, GYJF 19011 (LIP); trail head toward Roche Bateau, disturbed secondary rainforest, on bark, N3.620498, W53.199309, 22 June 2019, J. Fournier, GYJF 19209 (LIP) (immature). Stromata pulvinate to effused-pulvinate, 2 –10 mm long × 2–6 mm wide × 1–1.6 mm thick, coalescent into compound stromata up to 35 mm long × 13 mm wide with irregularly lobate contours, with unexposed to slightly exposed perithecial contours and abrupt margins; outermost coating dark brick (60) (Rayner 1970) with a faint vinaceous (57) tone, pruinose, wearing off on elevations and revealing a shiny blackish crust composed of waxy granules appearing light brown when bruised, honey (64) to cinnamon (62) when observed in water under the microscope, releasing sienna (8) KOH-extractable pigments within 1 min with a fugacious amber (47) halo, fading to light sienna with a marked vinaceous tone at 15 min, eventually pale greyish sepia (106) upon prolonged incubation; subperithecial tissue blackish, 0.3 – 0.8 mm thick, woody, with shiny black, carbonaceous, obliquely oriented strands. Perithecia lanceolate, 0.75 – 0.95 mm high × 0.2–0.25 mm diam. Ostioles umbilicate, most often inconspicuous. Paraphyses hyphal, thin-walled, remotely septate, 4 – 6 µm wide at base, tapering to 1.5 – 2 µm wide above asci, discretely embedded in mucilaginous material. Asci cylindrical, long-stipitate, originating from long ascogenous hyphae in unilateral spicate arrangement, with eight overlapping, uniseriately arranged ascospores, the sporebearing parts 60 – 69 × 6 –7.5 µm, the stipes fragile, 90 –150 µm long, with a discoid apical apparatus, apically convex with a sharp rim, 0.7– 0.9 × 2.1– 2.5 µm (av. 0.8 × 2.4 µm, n = 20), bluing in Melzer’s reagent. Ascospores (8.8 –) 9.4 –11.4(–11.8) × (4.5–)4.9–5.9(–6.3) µm, n = 120 (av. 10.4 × 5.4 µm), ellipsoid almost equilateral, with narrowly to less commonly broadly rounded ends, the lowermost ascospore in the ascus elongated with somewhat rhomboid ends, dark brown, with a wide straight germ slit c. 2/3 spore length with blurred contours, longitudinally to slightly obliquely oriented; perispore not dehiscent in 10 % KOH; epispore smooth Habitat & Distribution — On dead corticated branches in a tropical forest. Known only from Saül, French Guiana. Notes — Hypoxylon hepaticolor can be distinguished by its carbonaceous, effused-pulvinate stromata up to 1.6 mm thick with reddish brown surface and concolorous KOH-extractable pigments, lanceolate perithecia and ellipsoid-equilateral ascospores 10.4 × 5.4 µm on average, with a germ slit less than spore length and a perispore indehiscent in 10 % KOH. This combination of characters does not match any known Hypoxylon species. Carbonaceous stromata with lanceolate perithecia suggest possible affinities with the species recently reinstated in Pyrenopolyporus, accommodating taxa formerly placed in Hypoxylon characterised by peltate to discoid massive stromata over 2.5 mm thick and frequently irregularly shaped ascospores with a perispore indehiscent in 10 % KOH (Wendt et al. 2018). The stromata of the new species are not discoid or peltate and do not reach the thickness encountered in Pyrenopolyporus. Furthermore, affinities with this genus are not supported by our molecular results, which suggest affinities with Hypoxylon. Its three closest relatives in the LSU-based phylogenetic tree are notably different in having waxy-woody stromata with orange KOH-extractable pigments and ascospores with a spore-length germ slit and a perispore dehiscent in KOH. Hypoxylon fendleri MUCL54792 (KY610481) 64 Hypoxylon rickii MUCL53309 (KC968932) 98 Hypoxylon hepaticolor (MT799853) Hypoxylon lenormandii CBS119003 (KY610452) Daldinia vernicosa CBS119316 (KY610442) Daldinia petriniae MUCL49214 (KY610439) 0.005 Colour illustrations. French Guiana, Maripasoula, Saül, forest trail where the holotype was collected. Habit of coalescent stromata; stroma in vertical section; stromatal waxy granules observed in water; pigments released in 10 % KOH; long-stipitate ascus; ascospores; apical apparatus in Melzer’s reagent; germ slits on ascospores. Scale bars = 5 mm (stromata), 0.5 mm (section), 10 µm (granules, ascospores), 50 µm (ascus), 5 µm (apical apparatus, germ slits). Photos: Gilles Corriol (background) and Jacques Fournier. Maximum likelihood tree of LSU sequences generated using RAxML HPC2 (Stamatakis 2014) on the CIPRES v. 3.3 portal (Miller et al. 2010). Hypoxylon hepaticolor is in bold. RAxML bootstrap support values above 70 % are shown above the nodes and Bayesian posterior probability scores above 0.95 are shown as thickened branches. GenBank accession numbers for LSU sequences are given after taxon names. Andrew N. Miller, University of Illinois Urbana-Champaign, Illinois Natural History Survey, 1816 South Oak Street, Champaign, Illinois, 61820, USA; e-mail: amiller7@illinois.edu Jacques Fournier, Las Muros, 09420 Rimont, France; e-mail: jfournzeroneuf@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 358 Persoonia – Volume 45, 2020 Inocybe ionolepis 359 Fungal Planet description sheets Fungal Planet 1158 – 19 December 2020 Inocybe ionolepis Cullington & E. Larss., sp. nov. Etymology. Refers to the purple scales on the pileus. Classification — Inocybaceae, Agaricales, Agaricomycetes. Pileus 15 – 35 mm diam, campanulate to obtusely conical with an obtuse to broad umbo, slightly incurved margin, expanding with age, surface silky fibrillose with large flat depressed scales as young, dark purple and brown around the umbo and purple-lilac towards the margin, contrasting to the pale whitish subpellis trama. Fading with age to become pale greyish lilac at margin. Cortina pale greyish white with a violet tinge. Lamellae moderately crowded, interspersed with lamellulae, sinuate to emarginate, first pale grey with a lilac tint, later greyish brown. Stipe 35 – 45 × 3 – 5 mm, equal, equal to slightly bulbous, pale with greyish lilac tint, yellowish tint at the base, pruinose at apex to 1/3 of the stipe, in lower part fibrillose, covered with thin walled hyaline hyphae 5–12 µm wide. Smell slightly spermatic. Basidia clavate, 4-spored, (21.2 –)28.8(– 33) × (7–)8.6(–10.3) µm. Spores smooth ellipsoid to subamygdaliform with obtuse apex and distinct apiculus (8.6 –)9.5(–11.2) × (4.6 –)5.1(– 6.1) µm, Q = (1.82–)1.84(–1.96) (n = 85). Pleurocystidia (48 –) 5 4(–71) × (11–)14(–16) µm (n = 50), lageniform to subutriform, with pedicel, usually abundant with crystals at apex, walls 2 – 3.5 µm thick, mostly pale colourless. Cheilocystidia similar to pleurocystidia but shorter, (32 –) 40 (– 46) × (13 –)16(–18) µm (n = 20), lageniform to broadly utriform, walls 2.5 – 4.5 µm thick. Paracystidia hyaline, pyriform to clavate (14 –)17(– 22) × (8 –)9(–10) µm (n = 10). Caulocystidia at apex similar to pleurocystidia, abundant, with crystals, less so further down (40 –)50(–70) × (14 –)16(–18) µm (n = 10), fusiform to more cylindrical, caulocystidioid hairs thin-walled, sometimes septate, 40 –100 × 9 –12 µm, cauloparacystidia few. Pileipellis a compact interwoven cutis of cylindrical hyphae, thin-walled, smooth, hyaline (5 –)6 –10(–13) μm wide. Clamp connections present in all tissue. Ecology & Distribution — Associated with deciduous trees, Fagus sylvatica and Quercus robur. Basidiomata so far only known from England and Sweden, however ITS sequence data generated from soil samples show a wider distribution with occurrence in Iran, Estonia and Latvia. Notes — Inocybe ionolepis belongs in the I. geophylla group and the purple-lilac species surrounding I. lilacina (Matheny & Swenie 2018). The group is identified to host a high phylogenetic diversity and the name I. lilacina has been applied to many taxa in Europe (Ryberg et al. 2008). The group is still in need of further investigations with solid documentation of macromorphology and ecology. Inocybe ionolepis is characterised by having a pileus with a brown umbo and purple-lilac scales, and a pale yellowish tint at the stipe base. In dry condition the scales are distinctive, see the cap detail bottom right photo, but the scales colour fade and is affected after rain, and maybe also by late season fruiting. The young lamellae have a distinct lilac tone, that with age are becoming greyish brown with a pale lilac tint. Blast search of NCBIs GenBank nucleotide database and the UNITE database identified five additional ITS sequences of I. ionolepis generated from soil samples, suggesting the species to have a broad distribution range in Europe and Iran. In the phylogenetic analysis it comes out in a sister clade to I. lilacina 2 from Europe (Matheny & Swenie 2018). The two sequenced collections in the I. lilacina 2 clade originates from deciduous forests like I. ionolepis, while the two sequenced collections in I. lilacina 1 that comes out in a sister clade to I. sublilacina originates from coniferous forests. This suggests that there is an ecological differentiation within the group. Inocybe pallidicremea KY923033/KY923042 T 89 Inocybe pallidicremea KY990553/KY990505 78 Inocybe lilacina MH024860 T 65 Inocybe lilacina KY990525/KY990483 Inocybe lilacina 1 AM882869 100 Inocybe lilacina 1 AM882873 100 Inocybe sublilacina KY990561/JN974949 100 Inocybe sublilacina KY990562/KY990520 Inocybe ionocephala KY990549/JN974950 100 Inocybe ionocephala KY990551/KY990504 T Inocybe sp. FR852270 Iran 78 Inocybe sp. UDB017426 Iran Inocybe ionolepis MT909817 Sweden Typus. GreAt britAiN, England, Gloucestershire, Forest of Dean, near Acorn Patch, 21 Sept. 2017, P. Cullington, in deciduous forest on stony soil under Fagus sylvatica (holotype K(M)236689, isotype GB, ITS-LSU sequence GenBank MT909818, and MycoBank MB836902). Inocybe sp. UDB0268081 Latvia 100 Inocybe ionolepis MT909818 Great Britain T 97 Additional material examined. sweDeN, Gotland, Linde, Linde Prästänge, Forest meadow on calcareous ground, under Quercus robur, close to Corylus avellana and Betula pendula, 25 Oct. 2019, E. Larsson 279-19, GB-0207596, ITS-LSU sequence GenBank MT909817. Inocybe sp. UDB0106701 Estonia 100 Inocybe lilacina 2 AM882875 Inocybe lilacina 2 AM882874 Inocybe geophylla AM882870 100 Inocybe fuscicothurnata NR148184 T Inocybe fuscicothurnata MF487844 Inocybe flocculosa AM882891 20.0 Colour illustrations. Inocybe ionolepis habitat, forest meadow Linde prästänge, Gotland, Sweden. In situ basidiomata (GB-0207596); detail of pileus scales, cheilocystidia and basidiospores (holotype K(M)236689). Scale bars = 10 µm for spores, 20 µm cheilocystidia. Phylogram based on ITS and LSU sequence data showing the position of I. iololepis in the I. lilacina group. Bootstrap support values are indicated on branches and the sequence of the holotype is marked in bold. Multiple sequence alignments were carried out with MAFFT (https://mafft.cbrc.jp/ alignment/server/). The alignment was checked and adjusted manually, heuristic searches and bootstrap parsimony analyses were performed using PAUP v. 4.0b10 (Swofford 2003). Ellen Larsson, Biological and Environmental Sciences, University of Gothenburg, and Gothenburg Global Biodiversity Centre, Box 461, SE40530 Göteborg, Sweden; e-mail: ellen.larsson@bioenv.gu.se Penny Cullington, The Beeches, Pleck Lane, Kingston Blount, Oxfordshire, UK, OX39 4RU; e-mail: pennyculli@btinternet.com Kare Liimatainen, Jodrell Laboratory, Royal Botanic Gardens, Kew, Surrey TW9 3AB, UK; e-mail: K.Liimatainen@kew.org © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 360 Persoonia – Volume 45, 2020 Lactifluus lactiglaucus 361 Fungal Planet description sheets Fungal Planet 1159 – 19 December 2020 Lactifluus lactiglaucus P. Leonard & Dearnaley, sp. nov. Etymology. lactiglaucus means green milk and refers to the colour of the latex. Classification — Russulaceae, Agaricales, Agaricomycetes. Pileus centrally depressed to infundibuliform, 60–100 mm diam; surface dry, slightly velutinate, sometimes rugulose, usually with dirt adhering, azonate, white with some buff colouration at centre; margins in-rolled at first. Lamellae subdecurrent, crowded, anastomosing, off-white, very narrow (< 2 mm), turning slowly greenish on bruising and finally dirty brownish after some hours, lamellulae absent. Stipe cylindrical, 40 – 60 × 12–18 mm, glabrous, stout, very solid, white, green blotched if injured. Flesh white, thick, exuding a thick latex. Latex white, quickly turning greyish green to pistachio green (29D4-5; Kornerup & Wanscher 1978), turning orange or yellow with KOH. Smell of honey or baked bananas. Spore print white. Spores subglobose, a few broadly ellipsoid, 6.8 – 8.4 × 5.5 –7.4 µm, av. 7.3 ± 0.4 × 6.4 ± 0.5 µm, Q = 1.04 –1.3, Qav = 1.15 ± 0.06; ornamentation of low, slowly amyloid warts with fine lines joined to them like flagella, forming a partial reticulum; plage inamyloid, 2 μm (some spores remaining inamyloid at least in dried material). Basidia narrowly clavate, 45 – 50 × 6 – 8 µm, sterigmata 2 – 3 µm long, 2- and 4-spored basidia present. Pleurocystidia numerous, Lactifluus_albopicri_MEL2358392_AU Lactifluus_albopicri_MEL238012_AU Lactifluus_albopicri_AQ808493_AU Lactifluus_albopicri_MEL2044045_AU2 Lactifluus_albopicri_MEL2358395_AU Lactifluus_albopicri_MEL2238215_JET916_AU Lactifluus_albopicri_MEL2238214_AU Lactifluus_albopicri_MEL2358396_AU Lactifluus_albopicri_MEL2030453_AU Lactifluus_albopicri_MEL2037557_TWM1381_AU Lactifluus_albopicri_MEL2358394_TLe14_AU Lactifluus_albopicri_MEL2238277_LTe13_AU Lactifluus_albopicri_CD589_Au Lactifluus_albopicri_MEL2044045_AU Lactifluus_albopicri_MEL2297391_JET1203_AU Lactifluus_albopicri_MEL2371822_AU Lactifluus_albopicri_MEL2358395_TLe1123_AU Lactifluus_albopicri_MDBF12_18_AU Lactifluus_albopicri_DS69_Bald_Rock Lactifluus_aff._piperatus_KF220078_Thailand Lactifluus_aff._piperatus_KF220095_India Lactifluus_aff._piperatus_KF220105_Vietnam Lactifluus_aff._piperatus_KF220100_Thailand Lactifluus_piperatus_KF220037_France Lactarius_piperatus_KF220081_Denmark Lactarius_piperatus_KF220089_France Lactifluus_aff._piperatus_KF220102_Thailand Lactifluus_aff._piperatus_KF220112_Thailand thin-walled, narrowly clavate, 50 – 60 × 8 –10 µm, extending 10 –15 µm beyond basidia. Cheilocystidia numerous, similar to pleurocystidia, forming an almost sterile layer along the gill edge. Pileipellis an unusual type of ixocutis, resembling the hyphoepithelium illustrated (G on page 21) by HielmannClausen et al. (1998), hyphae in suprapellis only 3–4 µm wide. No lactifers seen in suprapellis. Habitat & Distribution — Gregarious in wet sclerophyll forest amongst leaf litter under Eucalyptus spp. So far only known from three sites in south east Queensland. Typus. AustrAliA, Queensland, Lamington National Park, 30 Mar. 2019, P. Leonard, (holotype PL640319 in BRI, ITS sequence GenBank MW007669, MycoBank MB837537). Additional materials examined. AustrAliA, Queensland, Bellthorpe, 21 Jan. 1985, T. Young, AQ646335 (BRI); New South Wales, Bald Rock National Park, 10 Apr. 2015, P. Leonard, PL630415 (BRI). Notes — This robust white fungus with hot peppery milk that turns pistachio green should be readily recognised in the field, yet it is only known from three collections. The earliest collection was identified as L. pergamenus, a synonym for the European species L. glaucescens. The European species is found in deciduous forests on calcareous soils and is said to be rather rare despite being reported from Northern Europe, North America and Japan. The Queensland collections are distinct, being found with Eucalyptus s.lat. in wet sclerophyll forests. Morphologically they are distinguished by more abundant milk that is almost immediately green and microscopically by the spores that are more globose than the European species. Its separation from the European collections is supported by our molecular analysis that places it in the same clade as the recently published L. austropiperatus and L. albopicrus (Crous et al. 2020a). There appear to be at least four Lactifluus species in section Piperates in Australia. They all have predominantly white fruiting bodies, crowded gills, hot to acrid tasting latex, and spores with low (< 0.5 µm) ornamentation. Lactifluus lactiglaucus is the most readily recognised on account of its green latex. Lactifluus_piperatus_KF220287_Belgium Lactifluus_austropiperatus_PERTH07550324_AU Lactifluus_austropiperatus_MEL2202701_AU Lactifluus_austropiperatus_AQ808481_AU Lactifluus_austropiperatus_Thiele2074_TLe1124_AU Lactifluus_dwaliensis_KF220111_Thailand Lactifluus_austropiperatus_ALV18132_Linda_Garrett Lactifluus_austropiperatus_ALV20037_Lamington Lactifluus_lactiglaucus_ALV18134_Bald_Rock Lactifluus_lactiglaucus_ALV20260_Lamington Lactifluus_aff._glaucescens_KF220049_USA Lactifluus_lactiglaucus_FG016IT_AU Lactifluus_leucophaeus_KF220058_PNG Lactifluus_leucophaeus_KF220058_Indonesia Lactifluus_leucophaeus_KF220059_Thailand Lactifluus_leucophaeus_GU258299_PNG Lactifluus_aff._glaucescens_KF220045_N._America Lactifluus_aff._glaucescens_KF220047_N._America Lactifluus_glaucescens_GU258298_Thailand Lactifluus_aff._glaucescens_KF220044_N._America Lactifluus_glaucescens_KF220070_Belgium Lactifluus_aff._glaucescens_KF220054_Thailand Lactifluus_glaucescens_KF220062_France Lactifluus_glaucescens_KF220025_Germany Lactifluus_aff._glaucescens_KF220103_Vietnam Lactifluus_aff._glaucescens_KF220051_Thailand Russula_nigricans_AF418607 Phylogenetic tree: Maximum likelihood tree of the ITS-nrDNA for a selection of Lactifluus species, aligned using MUSCLE and constructed using MEGA X. 0.09 Colour illustrations. Wet sclerophyll forest in south-east Queensland. Lower right sporocarp (holotype); lower centre right abundant green latex; lower centre left subglobose spores with low ornamentation; lower left pileipellis. Scale bars = 10 µm. All photos © Patrick Leonard. Patrick Leonard, P.O. Box 1193, Buderim 4556 Queensland, Australia; e-mail: patbrenda.leonard@bigpond.com John D.W. Dearnaley, Centre for Crop Health, University of Southern Queensland, Toowoomba 4350, Queensland, Australia; e-mail: John.dearnaley@usq.edu.au © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 362 Persoonia – Volume 45, 2020 Lophotrichus medusoides 363 Fungal Planet description sheets Fungal Planet 1160 – 19 December 2020 Lophotrichus medusoides Calvert, McTaggart & R.G. Shivas, sp. nov. Etymology. Named for the resemblance of ascomata to Medusa of Greek mythology, a Gorgon described as a winged woman with living venomous snakes for hair. Classification — Microascaceae, Microascales, Sordariomycetes. Mycelium on potato-dextrose agar (PDA) smooth, branched, sub-hyaline to pale yellow, hyphae 2 – 4 μm diam. Ascomata perithecial, immersed or partly immersed, globose to subglobose, scattered, 250 – 400 μm diam, with beaks 120 – 200 × 30 – 40 μm, with peridium composed of dark brown textura angularis; ascomatal appendages numerous, dark flexuous, up to 1.5 mm long, abundant on the beak, curved and narrowed at the apex. Asci evanescent, broadly clavate, 30 – 36 × 20 – 24 μm, thin-walled, 8-spored. Ascospores ellipsoidal, 7.5 – 9.5 × 5.5–6.5 μm, apices rounded, yellowish brown, with a germ pore at each end; wall even, 1–1.5 μm wide, smooth, extruded in a mass loosely held by ostiolar appendages. Culture characteristics — On PDA after 2 wk in the dark at 23 °C colonies 3.5 cm diam, flat, with sparse aerial mycelium, cream white, ascomata concentrated toward margin; with irregular darkened patches after 3 wk; reverse cream white. Typus. AustrAliA, Queensland, Iron Range, Lockhart River Rd, isolated from stem tissue of Citrus garrawayi (Rutaceae), 19 July 2019, J. Calvert (holotype specimen BRIP 70690, includes the ex-type culture BRIP 70690, ITS and LSU sequences GenBank MT180536 and MT186160, MycoBank MB834991). Notes — Lophotrichus medusoides was isolated as an endophyte from stem tissue of Citrus garrawayi, which is an understory tree endemic to tropical rainforests in the Cape York region of northern Queensland. Species of Lophotrichus have been described from mammal dung and soil and are morphologically similar to Microascus, Pseudallescheria and Kernia (SandovalDenis et al. 2016). Other taxa in the Microascales have been reported in association with Citrus as either endophytes, e.g., Scedosporium, or pathogens, e.g., Ceratocystis (De Beer et al. 2014). A phylogenetic analysis of the ITS region showed Lophotrichus was monophyletic and that L. medusoides shared a most recent common ancestor with coprophilic taxa. It differs in morphology to L. fimeti, which lacks beaks on ascomata, to L. plumbescens, which has two types of ascomatal hairs, and to L. martinii, which has larger ascospores and shorter, wavy ascomatal hairs (Guarro et al. 2012). It differs from L. indicus, which has broad ascospores with obtuse ends and ascomata that have abundant terminal hairs (Saxena & Mukerji 1970). BLASTn results of the ITS sequence of L. medusoides indicated similarity to type sequences of L. fimeti (NR_154109; Identities = 480/485 (98.97 %), no gaps), and Enterocarpus grenotii (NR_159852; Identities = 502/530 (94.72 %), six gaps (1 %)). The LSU sequence shared 1 056/1 096 (96 %) sequence identities with Cephalotrichum purpureofuscum (GenBank MF041789). A network analysis of ITS sequences showed a difference of three parsimony-informative characters between L. medusoides and L. martinii (GenBank MH856648) and four between L. medusoides and L. fimeti (GenBank MF161105). Microascus longirostris CBS196.61 NR132945 99.99/100 Microascus trigonosporus CBS218.31 NR136145 99.93/100 Pseudallescheria ellipsoidea CBS418.73 NR130665 Pseudallescheria fusoidea CBS106.53 NR145213 99.99/100 99.72/100 Kernia pachypleura CBS776.70 NR160127 Kernia geniculotricha CBS599.68 MH859190 L. macrosporus L. cf. ampullus L. fimeti strain UT-Ps2 86.73/84 Lophotrichus cf. ampullus UAMH11809 KM580494 L. fimeti TYPE L. martinii 99.82/100 L. plumbescens Lophotrichus medusoides BRIP70960 MT180536 L. medusoides 80.97/82 Lophotrichus martinii CBS326.50 MH856648 L. fimeti strain UT-Ps1 Lophotrichus sp. 89.37/92 Lophotrichus plumbescens NBRC30864 LC146745 76.96/77 n=1 Lophotrichus macrosporus CBS379.78 MH861152 n=2 78.6/81 Lophotrichus fimeti CBS129.78 NR154109 0.05 Colour illustrations. Monsoon rainforest in the Iron Range, Cape York Peninsula, Far North Queensland, Australia. Ascomata showing long ascomatal hairs; textura angularis; ascospores from ruptured ascomata; 8-spored unitunicate ascus; hyphae. Scale bars = 1 mm (top left), 10 μm (all others). Mid-point rooted phylogram from a maximum likelihood search using IQTREE v. 1.3.11.1 (Nguyen et al. 2015) with 10 000 ultra-fast bootstraps (Minh et al. 2013), 10 000 replicates of an approximate likelihood ratio test (aLRT), and a best-fit model of evolution (command -m TEST). ITS sequences aligned with MAFFT in UGENE v. 1.30.0 (Okonechnikov et al. 2012). The aLRT and UFBootstrap values are indicated at nodes. Minimum spanning network of all available Lophotrichus ITS sequences generated using POPART v. 1.7 (Leigh & Bryant 2015); hashes indicate number of parsimony informative characters between taxa. Jed Calvert, Alistair R. McTaggart, Lília Costa Carvalhais, André Drenth & Roger G. Shivas, Queensland Alliance for Agriculture and Food Innovation (QAAFI), University of Queensland, Ecosciences Precinct, Level 2C East, GPO Box 267, Brisbane 4001, Queensland, Australia; e-mail: J.Calvert@uq.net.au, A.Mctaggart@uq.edu.au, L.Carvalhais@uq.edu.au, A.Drenth@uq.edu.au & R.Shivas@uq.edu.au © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 364 Persoonia – Volume 45, 2020 Melanoleuca dominicana 365 Fungal Planet description sheets Fungal Planet 1161 – 19 December 2020 Melanoleuca dominicana Angelini, Para & Vizzini, sp. nov. Etymology. The name dominicana (Spanish) refers to the occurrence of the species in the Dominican Republic. Classification — Incertae sedis in the Pluteineae, Agaricales, Agaricomycetes. Pileus 4 – 5 cm diam, applanate, depressed with an umbilicate centre, rarely with a large and low umbo; pileus surface smooth, opaque, always very dark in the centre, brownish, up to blackish brown, otherwise ochre-brown, grey-brownish, also ash-grey. Lamellae medium crowded, with numerous lamellulae (l = 1– 3) of various lengths emarginated with long decurrent tooth, straight, white. Stipe 3.5 – 4 × 0.5 –1 cm, central, cylindrical, enlarged at the apex, clavate at the base, longitudinally fibrillose, from brown to dirty greyish brown, blackening at the base. Context white brownish in the pileus, brown in the stipe, brown blackish in the stipe base. Odour and taste not distinctive. Spores 5.8–7.8 × 4.8–6 µm (av. 7 × 5.3 µm, Q = 1.16–1.51, Qm = 1.32), subglobose, hyaline, warty; warts isolated rarely with thin ridges, with evident suprapicular zone, amyloid. Basidia 2–4-spored, 31.5–36 × 7.2–9.6 µm, clavate, with pedunculate fusiform base. Cheilocystidia 38.4 – 48 × 4.8 – 9.6 µm, very numerous, mainly nettle-hair shaped (urticoid) to fusoid with a transversal septum and crystals at the apex (exscissa-type). Pleurocystidia not observed. Paracystidia very numerous, cylindroid-clavate to irregularly clavate. Pileipellis a cutis with up to 4 µm wide hyphae, confusedly intertwined with few emerging elements. Pileocystidia not observed. Stipitipellis a cutis with parallel hyphae from which scattered cauloparacystidia emerge; outermost hyphae cylindrical, up to 170 µm long and 3.5 µm wide, in the inner layer cylindrical to allantoid, up to 48 µm long ad 8.5 µm wide. Caulocystidia not observed. Cauloparacystidia cylindroid to flexuous, cylindroid-clavate, sometimes bifurcate, 21.5 – 36 × 3.4 –7.2 µm. Lamellar trama regular, with parallel, slightly intertwined hyphae, 5 –7 μm wide. Clamp connections absent in all tissues. Habitat & Distribution — Growing solitary on tropical forest litter with both deciduous and coniferous trees, from sea level to the mountains. Uncommon in the studied area. So far known only from the Dominican Republic. Additional materials examined. DomiNicAN republic, Puerto Plata, Sosua, one basidiome collected on litter of a heavily anthropized humid woodland of deciduous trees, a few km from the sea, 29 Nov. 2013, C. Angelini JBSD130781; ibid., 30 Nov. 2013, C. Angelini, JBSD130780, ITS sequence GenBank MT991406. Notes — The new species belongs in subg. Urticocystis. The two collections of Melanoleuca dominicana clustered in a strongly supported clade (MLB = 100) sister to M. jaliscoensis and M. longisterigma clade but without support. Melanoleuca dominicana is well differentiated from the other Melanoleuca species described in literature, based on morphological and/or molecular characteristics. Melanoleuca tucumanensis, M. tucumanensis var. colorata and M. tucumanensis var. striata from Argentina (Singer & Digilio 1951, Raithelhuber 1974) have larger spores (7.5 –10.3 × 6.2 –7.5 µm, Singer & Digilio 1951; 7.2 – 9.6 × 4.8 –7.2 µm, pers. obs.). Despite several attempts, it was not possible to sequence neither the type nor other available collections of these three latter taxa. Melanoleuca jaliscoensis from Mexico differs from M. dominicana by its larger pileus (6.5–10 cm broad) and presence of pleurocystidia (Sánchez-García et al. 2013). Melanoleuca longisterigma from Mexico is distinguished by up to 10 µm long spores, a relevant percentage of mono- to bisporic basidia with long sterigmata and cylindrical to fusoid non-septate cheilocystidia without apical crystals (Sánchez-García et al. 2013). Melanoleuca yucatanensis from Mexico has pleurocystidia and shows shorter spores, (5.2 –)6 –7 µm long (Bon 1984). Typus. DomiNicAN republic, La Vega, Jarabacoa, two basidiomes on litter from a tropical mountain forest, with both broad-leaved and coniferous trees (Pinus occidentalis), 6 Dec. 2014, C. Angelini (holotype JBSD130779, ITS sequence GenBank MT991407, MycoBank MB837378). 100 Colour illustrations. Dominican Republic, La Vega, Jarabacoa, Pinus occidentalis forest, where the holotype specimen was collected. Melanoleuca dominicana basidiomata in field (holotype JBSD130779); fresh basidiomata after being collected (JBSD130780); fresh pileus (JBSD130781); lamellae attachment detail; cheilocystidia and spores; line drawings (spores, basidia, cheilocystidia, paracystidia, stipitipellis). Scale bars = 1 cm (basidiomes), 10 μm (cheilocystidia and spores, pictures). Melanoleuca brevipes JX429188 Melanoleuca humilis JX987317 Melanoleuca humilis KJ425531 Melanoleuca sublanipes JN616471 97 100 Melanoleuca sp. JN616479 Melanoleuca humilis MH930148 100 Melanoleuca rasilis JN616461 Melanoleuca sp. JF908344 99 Melanoleuca brevipes JF908352 Melanoleuca sp. JF908349 Melanoleuca longisterigma JX429211 88 100 Melanoleuca longisterigma JX429212 Melanoleuca jaliscoensis JX429222 100 Melanoleuca dominicana JBSD130780 MT991406 Melanoleuca dominicana JBSD130779 MT991407 Holotype Melanoleuca grammopodia JX429193 97 Melanoleuca grammopodia f. macrocarpa JF908350 Melanoleuca sp. JX429220 100 Melanoleuca herrerae JX429199 Melanoleuca herrerae JX429224 81 89 Melanoleuca exscissa JN616431 81 Melanoleuca diverticulata JN616429 Melanoleuca substrictipes JN616474 77 Melanoleuca strictipes JF908359 71 86 Melanoleuca iris JN616446 99 Melanoleuca exscissa JN616433 Melanoleuca substrictipes var. sarcophylla JN616475 Melanoleuca verrucipes JF908354 Melanoleuca decembris JN616426 100 Melanoleuca stridula JX429117 86 Melanoleuca pseudoluscina JF908346 100 Melanoleuca microcephala KX387848 Melanoleuca decembris JN616427 Melanoleuca pseudoluscina JN616458 100 Melanoleuca pseudoluscina JN616455 94 Melanoleuca paedida JN616452 Melanoleuca electropoda JN616430 Melanoleuca cognata JX429225 Melanoleuca paratristis JF908357 90 Melanoleuca striimarginata JN616468 Melanoleuca stridula JN616467 81 Melanoleuca graminicola JN616438 Melanoleuca sp. 'acystis group‘ JX429213 85 Melanoleuca angelesiana JN616420 Melanoleuca sp. JX429216 Melanoleuca robertiana JN616462 Melanoleuca polioleuca JX429195 Melanoleuca communis JX429205 Melanoleuca arcuata JX429186 91 77 75 0.03 Maximum-likelihood analysis of the nrITS region of Melanoleuca subg. Urticocystis species was performed with RAxML v. 8 (Stamatakis 2014) using the GTR+G model (1 000 bootstrap replicates). Only maximum-likelihood bootstrap support values ≥ 70 % are shown in the phylogenetic tree. The scale bar represents the number of nucleotide changes per site. Alfredo Vizzini & Francesco Dovana, Department of Life Sciences and Systems Biology, University of Torino, Viale P.A. Mattioli 25, I-10125 Torino, Italy; e-mail: alfredo.vizzini@unito.it & francescodovana@gmail.com Claudio Angelini, Herbario Jardín Botánico Nacional Dr. Rafael Ma. Moscoso, Santo Domingo, Dominican Republic and Via Cappuccini, 78/8 – 33170 Pordenone, Italy; e-mail: claudio_angelini@libero.it Roberto Para, Via Martiri di Via Fani 22, I-61024, Mombaroccio (PU), Italy; e-mail: r.para@alice.it © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 366 Persoonia – Volume 45, 2020 Moelleriella puertoricoensis 367 Fungal Planet description sheets Fungal Planet 1162 – 19 December 2020 Moelleriella puertoricoensis Mongkolsamrit, Noisripoom & Luangsa-ard, sp. nov. Etymology. Name refers to Puerto Rico, the location where this species was collected. Classification — Clavicipitaceae, Hypocreales, Sordariomycetes. Specimens found on the underside of dicotyledonous leaves of forest plants. Hosts are scale insect nymphs (Hemiptera). Stromata discoid, distinctly stud-shaped, up to 3 mm diam, and 1–2.5 mm high, pale yellow, base surrounded by a white membranous hypothallus. Conidiomata scattered around a narrow neck, with a pale yellow to yellow mass of conidia. Conidiogenous cells phialidic, aschersonia-like, cylindrical, straight, up to 55 µm long, 1– 2 µm wide, forming a compact layer. Conidia hyaline, fusoid, with acute ends, aseptate, (10–)11–12.5(–14) × 2–3 µm. Paraphyses absent. Hirsutella-like synasexual morph scattered on the upper surface of stroma, phialides with a long thin neck, entire phialides up to 25 µm, 3 – 5 µm wide, conidia globose, 4 – 5 µm diam. Sexual morph not observed. Culture characteristics — Colonies developed from germinating conidia. The conidia germinated within 24 h on potato dextrose agar (PDA). Colonies reaching a diam of 1 cm after 3 wk at 25 °C. Colonies compact with white mycelium, colonies reverse uncoloured. Conidia produced after 30 d, hyaline thinly spreading on colonies, fusoid, with acute ends, aseptate, 9–12 × 2– 3 µm. Typus. USA, puerto rico, Rio Abajo State Forest, on scale insect (Hemiptera) attached to underside of dicotyledonous leaves, 19 Jan. 2018, S. Mongkolsamrit, J.J. Luangsa-ard & S. Wongkanoun (holotype BBH43763, culture ex-type BCC88320, ITS, LSU and tef1 sequences GenBank MW115297, MN954683 and MN944389, MycoBank MB834780). Additional materials examined. USA, puerto rico, Rio Abajo State Forest, on scale insect (Hemiptera) attached to underside of dicotyledonous leaves, 19 Jan. 2018, S. Mongkolsamrit, J.J. Luangsa-ard & S. Wongkanoun, BBH43763 (BBC88321), ITS, LSU and tef1 sequences GenBank MW115298, MN954684 and MN944390; BBH43764 (BCC88322), ITS, LSU and tef1 sequences GenBank MW115299, MN954682 and MN944391. Notes — The gross macromorphology of the natural samples of M. puertoricoensis closely resembles the asexual morph of M. basicystis (Chaverri et al. 2008) and M. pongdueatensis (Li et al. 2016) that were found in Costa Rica and Thailand, respectively. These three species have discoid, distinctly stud-shaped stroma and conidiomata scattered around a narrow neck of the stroma, with a pale yellow to yellow mass of conidia. Conidia in M. puertoricoensis and M. pongdueatensis are fusoid with acute ends and the width of conidia are in the same range (10 –14 × 2 – 3 µm vs 9 –12.5 × 1.5 – 2.5 µm). Conidia of M. basicystis are ventricose with acute ends (11–15.5 × 3 – 5 µm) and wider than those reported in M. puertoricoensis and M. pongdueatensis. Moelleriella puertoricoensis and M. basicystis lack paraphyses while they are present in M. pongdueatensis, linear, filiform, up to 110 × 1– 2 µm. Additionally, M. puertoricoensis and M. pongdueatensis have a hirsutella-like synasexual morph scattered on the upper surface of the stroma, phialides with a long thin neck (25 × 5 μm vs 20 × 1– 2 µm), globose (4 – 5 µm) and citriform (2 – 3 × 1– 2.5 μm) conidia, respectively. The results of our molecular phylogenetic study strongly support and separate M. puertoricoensis from other known species. Moelleriella puertoricoensis is therefore proposed as a new species belonging to Moelleriella from the Neotropics. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Moelleriella basicystis (strain ROKI2770, GenBank EF190282.1; Identities = 546/594 (92 %), 21 gaps (3 %)), and Orbiocrella sp. (strain BCC33248, GenBank KJ138267.1; Identities = 300/349 (86 %), 22 gaps (6 %)). Closest hits using the LSU sequence had highest similarity to Moelleriella basicystis (strain F183147, GenBank EU392577.1; Identities = 276/279 (99 %), 1 gap (0 %)), and Moelleriella phyllogena (strain CUP 67340, GenBank AY518372.1; Identities = 276/279 (99 %), 1 gap (0 %)). Closest hits using the tef1 sequence are Moelleriella basicystis (strain F183147, GenBank EU392653.1; Identities = 856 /912 (94 %), no gaps (0 %)), and Moelleriella basicystis (strain P.C. 374, GenBank AY986928.1; Identities = 854/910 (94 %), no gaps (0 %)). Supplementary material Colour illustrations. Background photo of side of a trail in Rio Abajo State Forest; fungi on hosts, side view of stroma showing stud-shaped and conidiomata, conidiogenous cells, conidia, hirsutella-like on stroma, conidia, culture derived from conidia on PDA (sporulation present). Scale bars = 1 and 2 mm (stromata), 20 µm (phialides), 10 µm (conidia, hirsutella-like phialides and conidia), 3 mm (culture). FP1162-1 Phylogenetic reconstruction of M. puertoricoensis was done using a combined dataset comprising LSU and tef1 sequences. The data was analysed using Maximum parsimony (MP), Maximum likelihood (ML) and Bayesian inference. The MP analysis was conducted on the combined data set using PAUP v. 4.0b10 (Swofford 2002), adopting random addition sequences (100 replications), with gaps being treated as missing data. A bootstrap (BP) analysis was performed using the maximum parsimony criterion in 1 000 replications. The ML analysis was run with RAxML-VIHPC2 v. 8.2.12 (Stamatakis 2014) under a GTR model, with 1 000 bootstrap replicates. Bayesian phylogenetic inference was calculated with MrBayes v. 3.2.7a (Ronquist et al. 2012), with 5 M generations and under the same model. Numbers at the significant nodes represent MP bootstrap support values/RAxML bootstrap support values/Bayesian posterior probabilities (BPP) times 100. Thickened lines in the tree represent 99 –100 % bootstrap support values and 99 –100 BPP. FP1162-2 List of species and GenBank accession numbers of sequences used in this study. Suchada Mongkolsamrit, Wasana Noisripoom & Janet Jennifer Luangsa-ard, National Center for Genetic Engineering and Biotechnology (BIOTEC), 113 Thailand Science Park, Phahonyothin Rd., Khlong Nueng, Khlong Luang, Pathum Thani 12120, Thailand; e-mail: suchada@biotec.or.th, wasana.noi@biotec.or.th & jajen@biotec.or.th © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 368 Persoonia – Volume 45, 2020 Mycena pulchra 369 Fungal Planet description sheets Fungal Planet 1163 – 19 December 2020 Mycena pulchra P. Leonard, sp. nov. Etymology. pulchra means pretty and refers to the fungus in its setting on a paperbark tree. Classification — Mycenaceae, Agaricales, Agaricomycetes. Pileus convex with a central umbilicus, 15–40 mm diam, bright reddish pink (9A6, 11A8; Kornerup & Wanscher 1978), flamingo pink, fading to pale pink with age; margin ± smooth. Lamellae adnate or with a subdecurrent tooth, white, lamellulae arranged in two series alternating with lamellae, 16–18 lamellae reach the stipe. Stipe cylindrical to somewhat flattened, tough, centrally attached, curved, 15 – 30 × 1.5 – 4 mm; bright reddish pink, but paler than cap and white towards base; fruiting singly or in small caespitose groups. Flesh white, thin. Spore print white. Spores ellipsoid, 10.4 –14.6 × 5.7– 8.8 µm (av. 13 ± 1.23 × 6.7 ± 0.84, Q = 1.5–2.5, Qav = 1.96 ± 0.3), spore contents weakly amyloid with Melzer’s reagent. Basidia strongly clavate, 45–60 × 10–12 µm, 4-spored. Pleurocystidia clavate, 44 – 50 × 7.5 –11 µm, amyloid granular contents. Cheilocystidia numerous, forming an almost sterile edge to the gill, 40 –100 × 8 – 20 µm; ventricose or narrowly utriform. Pileipellis an irregular cutis, hyphae 7–12 µm, clamps absent. Habitat & Distribution — Growing in borer holes on the living trunks of the swamp paperbark, Melaleuca quinquinervia. Sporocarps found from just above flood level to about 2 m above ground level. Also seen on wounds where the tree has been damaged by storms or pruning. Appears to follow the distribution of Melaleuca quinquinervia with confirmed records from Eastern Australia and Western New Caledonia. Expected in New Guinea but there are no records thus far according to Maas Geesteranus & Horak (1995). Typus. AustrAliA, Queensland, Tewantin, Heritage Park, 13 Apr. 2008, J. Heavey (holotype PL58408, Brisbane, ITS sequence GenBank MT988148, MycoBank MB837369). Notes — This flamingo pink fungus is associated with wounds and borer holes in live paperbark trees. It grows on the tree trunk beneath the layers of bark and requires a wound or an insect hole in order to emerge and fruit. There are 16 collections under the name Mycena roseilignicola on I-Naturalist. Two are recorded on Melaleuca and appear to conform with M. pulchra. Six others are on dead wood and exhibit the striate cap that Corner (1994) describes. The eight other specimens either lack sufficient information to form a judgement or exhibit attachment via a distinct mycelial pad which is not a feature of M. pulchra nor mentioned by Corner (1994) for M. roseilignicola. Mycena haematopus MK979367 USA Mycena galericulata MH930217 Russia Mycena rosella HMAS 290152 MK966538 Cruentomycena viscidocruenta EU517518 TAS Roridomyces roridus MT153145 Nor Mycena pulchra PL58408 QLD Mycena roseilignicola MEL:2382631 NT Rickenella fibula MF319093 USA 0.04 Colour illustrations. Paperbark forest in south-east Queensland. Lower right fruiting body emerging from borer hole (holotype); lower centre fruiting body in tree wound; lower left pileipellis. All photos © Patrick Leonard. Scale bars = 20 µm. Maximum likelihood tree of the ITS-nrDNA for a selection of Mycena species, aligned using MUSCLE and constructed using MEGA X. Patrick Leonard, P.O. Box 1193, Buderim 4556 Queensland, Australia; e-mail: patbrenda.leonard@bigpond.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 370 Persoonia – Volume 45, 2020 Penicillium vallebormidaense 371 Fungal Planet description sheets Fungal Planet 1164 – 19 December 2020 Penicillium vallebormidaense Houbraken & Di Piazza, sp. nov. Etymology. Latin, name refers to Valle Bormida, the region from which the type specimen was collected. Classification — Aspergillaceae, Eurotiales, Eurotiomycetes. Conidiophores monoverticillate; stipes non-vesiculate, smooth, short, 13 – 25(– 35) × 1.5 – 2.5 µm; phialides ampulliform, 3 – 6 per conidiophore, 7–8.5(–10) × 2–2.5(–3) µm. Conidia smooth, globose to subglobose, 2 – 2.5(– 3) µm. Ascomata or sclerotia not observed. Culture characteristics (25 °C, 7 d) — Czapek yeast extract agar (CYA): Colonies non-sulcate, elevated in centre; margin slightly irregular; mycelium pale yellow; texture velvety; sporulation absent; soluble pigments brown, moderately produced; exudates absent; reverse brown. Malt extract agar (MEA): Colonies non-sulcate, slightly elevated in centre; margin slightly irregular; mycelium pale yellow; texture floccose; sporulation poor in centre, profuse in a ring between centre and edge, absent at edge; soluble pigments absent; exudates absent; conidia en masse pale grey green; reverse brown, pale brown at edge. Yeast extract sucrose agar (YES): Colonies randomly sulcate (radial and concentric), slightly elevated; margins slightly irregular; mycelium pale yellow; texture floccose in centre, velvety at edge; sporulation absent; soluble pigment absent; exudates absent; reverse pale brown. Dichloran 18 % glycerol agar (DG18): Colonies non-sulcate, plane, raised at the centre; margins entire; mycelium pale yellow in centre, white at edge; texture velvety; sporulation absent; soluble pigments absent; exudates absent; reverse pale yellow-brown. Oatmeal agar (OA): Colonies non-sulcate, plane, low; margins slightly irregular; mycelium yellow; texture velvety; sporulation absent; soluble pigments present, pale brown, poorly produced; exudates absent. Creatine agar (CREA): poor growth, acid production absent, base production absent. Colony diam, after 7 d, in mm — CYA 18–22; CYA 30 °C 22–26; CYA 37 °C 16–20; MEA 18 – 22; DG18 20 – 23; YES 21– 25; OA 18 – 22; CREA 9–11. Based on the phylogenetic analysis, P. vallebormidaense belongs to series Erubescentia of section Exilicaulis. Penicillium vallebormidaense grows rather slowly on CYA, MEA, DG18 and YES, is able to grow at 37 °C and produces short, monoverticillate conidiophores. These features are shared with many other species in series Erubescentia (Houbraken et al. 2020), confirming the results of the phylogenetic analysis. The new species is phylogenetically most closely related to NRRL 739, the ex-type of P. vinaceum. The sequence similarity scores with this strain are: BenA 95.0 % (identities = 459/478), CaM 90.6 % (identities = 444/490) and RPB2 93.8 % (identities = 837/892). Penicillium vinaceum is characterised by the (copious) production of ruby or vinaceous exudates on CYA and other media. In contrast, no exudate production is observed in P. vallebormidaense. Furthermore, the mycelium of P. vallebormidaense is pale yellow coloured, while the mycelium in P. vinaceum is white (Pitt 1980). Series Erubescentia P. guttulosum P. pimiteouiense P. rubidurum P. vallebormidaense sp. nov. P. vinaceum P. labradorum P. parvum P. menonorum P. nepalense P. catenatum P. striatisporum P. canis Typus. itAly, Savona, Valle Bormida, Ferrania (Cairo Montenotte), from compost 18 d in maturation, 26 June 2018, S. Di Piazza (holotype CBS H-24527, culture ex-type CBS 147064 = DTO 402-H5; ITS, LSU, BenA, CaM and RPB2 sequences GenBank MT316359, MW092765, MW115862, MW115863 and MW115864; MycoBank MB837659). P. erubescens 71 P. hermansii 87 P. laeve P. ovatum Notes — A BLAST search of BenA, CaM and RPB2 sequences of P. vallebormidaense CBS 147064 against an in-house reference sequence database containing data of all accepted Penicillium species (Houbraken et al. 2020), did not retrieved any high similarity hits. A homology search with the ITS sequence retrieved P. pimiteouiense (99.2 %), P. guttulosum (99.0 %) and P. vinaceum (98.9%) as most similar species. P. dimorphosporum 77 P. parvofructum P. corylophilum P. lapidosum P. citreonigrum P. restrictum P. alutaceum 0.01 P. inusitatum P. fractum Colour illustrations. Compost pile during ripening process. Colonies (7 d, 25 °C), left to right, first row: CYA, MEA, second row: CYA reverse, YES observe; conidiophores; conidia. Scale bars = 10 µm. Maximum likelihood tree of Penicillium strains belonging to section Exilicaulis series Erubescentia based on 1 871 aligned nucleotides (combined BenA, CaM and RPB2 sequences). Strain and GenBank accession numbers used in the analysis can be found in Houbraken et al. (2020). Analysis performed using RAxML v. 8.2.12 (Stamatakis 2014). Bootstrap 1 000 re-samplings; only bootstrap support values above 70 % are presented at the nodes and branches of > 95 % are thickened. Penicillium fractum and P. inusitatum were used as outgroup. The scale bar indicates the number of substitutions per site. Simone Di Piazza, University of Genoa, Department of Earth, Environmental and Life Science, Laboratory of Mycology, Corso Europa 26, 16132 Genoa, Italy; e-mail: simone.dipiazza@unige.it Jos Houbraken, Westerdijk Fungal Biodiversity Institute, Uppsalalaan 8, 3584 CT Utrecht, The Netherlands; e-mail: j.houbraken@wi.knaw.nl © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 372 Persoonia – Volume 45, 2020 Penicillium saanichanum 373 Fungal Planet description sheets Fungal Planet 1165 – 19 December 2020 Penicillium saanichanum Visagie, Assabgui & Seifert, sp. nov. Typus. cANADA, North Saanich, from house dust, May 2017, coll. B. Kendrick, isol. C.M. Visagie (holotype DAOM 745787, cultures ex-type DAOMC 251850 = KAS 6184; LSU, ITS, BenA, CaM and RPB2 sequences GenBank MN807447, KY469059, KY469096, KY469020, MN795070, MycoBank MB835962). Etymology. Latin, saanichanum, named after Saanich, the municipality where the noted Canadian mycologist Bryce Kendrick collected the house dust sample that this species was isolated from. Classification — Aspergillaceae, Eurotiales, Eurotiomycetes. Notes — A BLAST search of our BenA sequence against a locally curated reference dataset placed the new species in section Cinnamopurpurea series Idahoensia (Visagie et al. 2014, Houbraken et al. 2020). Penicillium saanichanum is characterised by restricted growth and monoverticillate conidiophores, characters typical of species classified in section Cinnamopurpurea. Morphologically and phylogenetically it is most similar to P. idahoense. However, the new species is morphologically distinct from P. idahoense based on its generally more restricted growth on most agar media, its red soluble pigments produced on CYA, and the absence of sclerotia (Paden 1971, Pitt 1980). Conidiophores monoverticillate and loosely divaricate; stipes smooth, 19–60 × 2–3 μm; vesicles 3–4.5 μm; branches 16–23 μm; phialides ampulliform, 5 –12 per metula, 7.5 –11 × 2.5 – 3.5 μm (8.6 ± 1.1 × 2.6 ± 0.2); conidia smooth, subglobose to globose, 2 – 3 × 2 – 3 μm (2.7 ± 0.2 × 2.5 ± 0.1), av. width/length = 0.92, n = 72. Culture characteristics (25 °C, 7 d) — On Czapek yeast autolysate agar (CYA): Colonies moderately deep, sunken in centre, slightly sulcate; margins low, narrow, entire; mycelia white; texture velutinous; sporulation moderate, conidia en masse greenish grey to dull green (27C2 – D3 – 4; colour codes based on Kornerup & Wanscher (1967)); soluble pigments red, inconspicuous; exudates absent; reverse dark ruby (12F8). On Blakeslee’s malt extract agar (MEA): Colonies moderately deep, planar; margins low, narrow, entire; mycelia white; texture velutinous; sporulation moderately dense, conidia en masse dull green to greyish green (26D3 – 4 – E5); soluble pigments red, inconspicuous; exudates absent; reverse violet brown to dark ruby (10E6–12F8). On 20 % sucrose CYA (CYA20S): Colonies with conidia en masse greyish green (26D5 – E5), otherwise similar to CYA. On 20 % sucrose MEA (MEA20S): Colonies less dense than those on MEA, lacking soluble pigment and red reverse colour, otherwise similar to MEA. On dichloran 18 % glycerol agar (DG18): Colonies similar to those on MEA. On yeast extract sucrose agar (YES): Colonies similar to those on MEA. On creatine sucrose agar (CREA): Growth good, no acid produced. Colony diam (in mm): CYA 9–11; CYA37C no growth; CYA20S 10 –12; MEA 7– 8; MEA20S 9 –10; DG18 9 –12; YES 12 –15; OA 4– 5; MY1012 7– 8; MY50G 3 – 4; CREA 5– 6. P. saanichanum DAOMC251850T KY469059 KY469096 KY469020 MN795070 P. colei NRRL13013T KF932958 KF932926 KF932942 KF932996 P. ellipsoideosporum CBS112493T JX012224 JQ965104 AY678559 JN121427 1/100 1/93 P. monsserratidens NRRL35884 KF932967 KF932935 KF932952 KF933006 1/99 P. monsserratidens NRRL35840T KF932962 KF932930 KF932947 KF933001 P. monsserratidens NRRL62003 KF932973 KF932940 KF932957 KF933010 P. idahoense CBS341.68T KC411747 EF626953 EF626954 JN121499 1/100 P. cvjetkovicii NRRL35841T KF932963 KF932931 KF932948 KF933002 P. infrapurpureum CBS138219T KJ775679 KJ775172 KJ775406 1/96 P. minnesotense NRRL66823T MK791277 MK803429 MK803430 MK796158 1/95 P. salmoniflumine NRRL35837T KF932960 KF932928 KF932945 KF932999 1/96 1/94 P. salmoniflumine NRRL58001 KF932969 KF932954 KF933008 1/84 P. monsgalena NRRL22302T KF932959 KF932927 KF932943 KF932997 1/99 P. lemhiflumine NRRL35843T KF932964 KF932932 KF932949 KF933003 P. fluviserpens NRRL35838T KF932961 KF932929 KF932946 KF933000 P. malacaense CBS160.81T EU427300 EU427268 KJ866997 JN406626 1/99 P. cinnamopurpureum CBS429.65T EF626950 EF626948 EF626949 JN406533 P. parvulum CBS132825T EF422845 EF506218 EF506225 P. gravinicasei NRRL66733T MG600580 MG600565 MG600570 MG600575 1/91 P. incoloratum CBS101753T KJ834508 KJ834457 KJ866984 JN406651 1/100 P. shennangjianum CBS228.89T KC411705 KJ834491 AY678561 JN121458 P. nodulum CBS227.89T KC411703 KJ834475 KJ867003 JN406603 1/100 P. jiangxiense AS3.6521T KJ890411 KJ890409 KJ890407 series Jiangxiensia P. pusillum CBS312.63T EF626951 KF932925 KF932941 KF932995 P. lunae PPRI25881T MK450725 MK451088 MK451660 MK450863 1/99 1/98 1/- series Idahoensia 1/0.97/- series Cinnamopurpurea x8 0.05 Colour illustrations. Bryce Kendrick’s home laboratory. Colonies on CYA and MEA; conidiophores; conidia. Scale bars = 10 µm. 1/100 Combined phylogeny of Penicillium section Cinnamopurpurea based on ITS, BenA, CaM and RPB2. Aligned data sets (MAFFT v. 7.450; Katoh & Standley 2013) were analysed using Maximum Likelihood (IQ-tree v. 1.6.12; Nguyen et al. 2015) and Bayesian Inference (MrBayes v. 3.2.7a; Ronquist et al. 2012). Bootstrap support values (≥ 80 %) and posterior probabilities (≥ 0.95) are given above branches. The new species is indicated by bold text, T = ex-type strain and GenBank accession numbers are shown in a smaller font next to the culture accession number (ITS = green, BenA = blue, CaM = red, RPB2 = purple). The tree is rooted to P. lunae. Cobus M. Visagie, Department of Biochemistry, Genetics and Microbiology, Forestry and Agricultural Biotechnology Institute (FABI), University of Pretoria, Pretoria, South Africa; e-mail: cobus.visagie@fabi.up.ac.za Rafik Assabgui, Biodiversity (Mycology), Agriculture and Agri-Food Canada, Ottawa, ON K1A0C6, Canada; e-mail: rafik.assabgui@canada.ca Keith A Seifert, Biodiversity (Mycology), Agriculture and Agri-Food Canada, Ottawa, ON K1A0C6, Canada / Department of Biology, Carleton University, 1125 Colonel By Drive, Ottawa, ON K1S 5B6, Canada; e-mail: keith.seifert@carleton.ca © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 374 Persoonia – Volume 45, 2020 Phialocephala melitaea 375 Fungal Planet description sheets Fungal Planet 1166 – 19 December 2020 Phialocephala melitaea Matočec, I. Kušan, Pošta, Tkalčec & Mešić, sp. nov. Etymology. Lat. melitaeus - which refers to Mljet (lat. Melita, ex C. Plinius Secundus), island in the Adriatic Sea. Classification — Mollisiaceae, Helotiales, Leotiomycetes. Ascomata apothecial, shallowly cupulate and with basal depression when young, plate shaped to plane when fully mature, superficial, sessile, ± circular to much irregular from the top view, *0.8 –1.4(– 2) mm diam, gregarious but mutually distanced. Hymenium, when in fresh state, in the central part greyish to pale lead-grey with bluish tones, perimarginal area beige to pale grey, not wrinkled but finely pruinose; margin ± irregular, lobed and wavy in fully mature apothecia, entire, ± sharp, white, subglabrous; excipular surface in upper part whitish, lower part brownish, slightly roughened. Subicular hyphae macroscopically not discernible. Hymenium *70 – 85 µm thick. Asci cylindrical with conical-subtruncate to rounded apex, *60–82 × (6.4–)6.7–8.1 µm, pars sporifera *19–25.5 µm, 8-spored, in living state protruding above paraphyses up to 18 µm, base cylindrical-truncate, arising from croziers, in Lugol’s solution (IKI) apical ring of medium amyloidity (2bb), Calycinatype. Ascospores piscioid to subscutuliform, with tapered to somewhat rounded poles, sometimes slightly bent, aseptate, *(7.8 –)8.2– 9.8–11.5(–12.6) × (2.4 –) 2.6 – 2.9– 3.2(– 3.4) µm, *Q = 2.7–3.4 – 4.1(– 4.3), hyaline, smooth, uninucleate, nearly eguttulate or with several dispersed tiny lipid bodies (LBs), *0.4 – 0.7 µm diam, freshly ejected without sheath remnants, biseriate inside *asci; in IKI unstained, nucleus slightly contrasted. Paraphyses cylindrical-obtuse, rarely apically clavate, apical cell *27– 63 × 3.4 – 5 µm, straight, simple, thin-walled, *containing few globose or obloid rarely elongated, moderately refractive and hyaline semi-resistant vacuolar bodies (SVB), with age become yellowish; in †KOH without yellow reaction; in *IKI SVBs golden yellow; in *brilliant cresyl blue (CRB) violet blue and after addition of †KOH light ruby red. Subhymenium *15 – 21 µm thick at the middle flank, hyaline, composed of densely packed epidermoid cells, *3.2 – 5.6 µm wide. Medullary excipulum *18 – 28 µm thick at the middle flank, reaching *36 µm in the lower flank, subhyaline to hyaline, stretches as a continuous layer towards the margin, composed of textura prismatica, cells *12 – 28 × 3 –10.5 µm, thin-walled, slightly gelatinised (purple in CRB), devoid of crystals. Ectal excipulum *28 – 40 µm thick at the middle flank, reaching *58 µm in the basal part, ochre brown to greyish brown, composed of textura angularis with elongated elements, cells *9 – 25.6 × 5.6 –14 µm, walls *0.7– 0.9 µm thick, upper flank hyaline and indistinguishable from medulla, terminal cells contain single, hyaline and globose SVB, lower flank rich in dark brown intercellular pigment, cell walls brown, thickened, *0.8 –1.2 µm. Marginal tissue completely hyaline, *21–25 µm thick, composed of textura porrecta-prismatica, cells *7.5 –14.6 × 2.3 – 5.7 µm, Colour illustrations. Croatia, Mljet Island, near Prožurska blatina (type locality). Apothecia; 14-d-old colonies on MEA, colony centre with exudates; asci and croziers in *H2O, ascus in *IKI; ascospores in *H2O; paraphyses in *H2O, *IKI and *CRB; phialides with collarettes, conidiophores (two pictures); hyphal loops, colony hyphae bearing exudates, apothecial subicular hyphae with gel envelope, colony hyphae with plaques and pigmented granules, vertical median section of the apothecium. Scale bars = 10 mm (colony), 1 mm (apothecia, colony centre), 50 µm (apothecial anatomy), 10 µm (microscopic elements). thin-walled, outermost cells cylindrical-clavate, each containing single globose hyaline SVB. Subicular hyphae sparse, confined to an apothecial base and lower flank, flexuous, greyish-brown, smooth, cells *19 – 23 × 3.3 – 5 µm, walls *0.6 – 0.8 µm thick, enveloped in gel up to *1.8 µm thick. Asterisk (*) denotes living and cross (†) dead state. Ascus amyloidity is termed after Baral (1987) and spore shape after Kušan et al. (2014). Colonies after 14 d in the dark at 24 °C on 3 % malt extract agar (MEA) 32–34 mm diam, centrally pronouncedly papillate, with woolly aerial hyphae; margin radially diffuse, hyaline; surface whitish grey near the margin, pale clay buff towards the centre, clay pink in the centre; reverse dark grey to blackish sepia. Exudates present in the central part, droplets honey brown. Conidiophores developed only after 5 mo, mycelium consisting of hyaline to subhyaline hyphae only in colonial margin, others brown-walled, smooth, septate, branched, *2.2 – 4.2 µm wide, wall *0.3 – 0.6 µm, often producing loops or covered with large tuberculate exudates, *1.5–3.8 µm high, some hyphae produce large flat subhyaline plaques under which granular brown pigment develops. Conidiophores micronematous, crawling or erect, greyish brown, smooth, cylindrical, with slightly thickened walls, conidiophore stem *9 –13 × 4 – 4.8 µm, 0 –1-septate, rami richly and tightly branched. Conidiogenous cells phialidic, terminal, conoid, ventricose or ampulliform, *7.8 –10.5 × 3 – 4.1 µm, collarettes flaring, *1.5–2.2 × 1.4–2.1 µm, hyaline. Conidia hyaline, dimorphic, primary conidia oblong to narrowly ellipsoid, aseptate, *2.9 – 4.3 × 1.6 –1.7 µm, secondary conidia globose to subglobose, aseptate, base often subtruncate, *1.8–2.3 µm. Distribution & Habitat — Known so far only from the type locality on the island of Mljet, Croatia. Type collection was found on a branch of Pinus halepensis lying in thick litter, in the evergreen thermo-Mediterranean forest. Typus. croAtiA, Dubrovnik-Neretva County, Island of Mljet, Prožura, near Prožurska blatina, 5 m asl, N42°43'57" E17°38'41"; on fallen decorticated main branch of Pinus halepensis, majority of the apothecia were growing on old fruitbody of Phellinus sp., only partially directly on wood, in a forest of P. halepensis, Viburnum tinus, Myrtus communis, Arbutus × andrachnoides, Coronilla sp. and Laurus nobilis, 16 Mar. 2020, I. Kušan & M. Pucar (holotype CNF 2/11027, ex-type culture CBS 147182, ITS and LSU sequences GenBank MT957536 and MT957586, MycoBank MB836891). Notes — Tanney & Seifert (2020) assigned species of true Phialocephala to a genus separated from Mollisia s.str., a view accepted here. The most closely related species Phialocephala biguttulata differs by ascospores containing two large polar guttules, paraphysis apical cells largely occupied by elongated refractive VBs, excipular tissue stained deep green by KOH and finally not producing conidiogenous structures in axenic culture. There are two other species, P. cladophialophoroides and P. aylmerensis, members of the same clade. The apothecial morph is not known for the first species whereas it produces acropetal conidial moniliform cell chains (cf. Crous et al. 2017b), and the latter differs in having more stout and (text continues on Supplementary material page FP1166) Supplementary material FP1166 Phylogenetic tree obtained from maximum likelihood analysis based on ITS sequences of Phialocephala melitaea and related species. Neven Matočec, Ivana Kušan, Ana Pošta, Zdenko Tkalčec & Armin Mešić, Ruđer Bošković Institute, Bijenička cesta 54, HR-10000 Zagreb, Croatia; e-mail: nmatocec@irb.hr, ikusan@irb.hr, aposta@irb.hr, ztkalcec@irb.hr & amesic@irb.hr © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 376 Persoonia – Volume 45, 2020 Phytophthora aquae-cooljarloo 377 Fungal Planet description sheets Fungal Planet 1167 – 19 December 2020 Phytophthora aquae-cooljarloo R. Mostowfizadeh-Ghalamfarsa & T.I. Burgess, sp. nov. Etymology. Named for the association of this species with water at the place Cooljarloo. Classification — Peronosporaceae, Peronosporidae, Oomycota. Sporangia produced on V8 agar (V8A) and carrot agar (CA) flooded with both distilled water and non-sterile soil extract; terminal, non-papillate, mostly ellipsoid to ovoid and limoniform, sometime obovoid; 66 ± 11.2 × 37 ± 4.5 μm (overall range 38 –101 × 29 – 54 μm), length/breadth ratio of 1.9 ± 0.2. Sporangial proliferation in chains of internally proliferating sporangia, both nested and extended. Hyphal swellings absent. Chlamydospores common, globose, thin-walled, 9 – (20 ± 5) – 33 µm. Gametangia were produced in single cultures (homothallic). Oogonia, smooth-walled, globose, golden to brown 34 ± 4.5 μm (isolates ranged from 23–47 μm). Oospores were aplerotic, with an average of 30 ± 4 μm (isolate means 19 to 41 μm). Oospore walls were av. 3.2 ± 0.9 μm, oospore wall index 0.5. Antheridia were paragynous, monoclinous, spherical to ellipsoidal, ranged from 5 to 17 μm in length (av. 12 ± 1.7 μm) and 5 to 14 μm in breadth (av. 9 ± 1.7 μm). Hyphae were hyaline, normally not septate, 4–5 μm wide. Minimum, optimum and maximum temperatures for growth were 4 °C, 30 °C and 35 °C, respectively. Radial growth rate on CA in the dark at 30 °C was 3.9 ± 1.3 mm/d. Culture characteristics — The colony patterns on all media were uniform with the exception of potato dextrose agar (PDA) which produced rose-shaped pattern in some isolates. Aerial mycelia were observed on some colonies specially on PDA. Notes — Isolates of Phytophthora aquae-cooljarloo constitute a well-supported monophyletic group sharing a common ancestor with P. gemini (Man in ’t Veld et al. 2011). These species together with P. humicola (Ko & Ann 1985), P. inundata (Brasier et al. 2003), P. condilina (Burgess et al. 2018), P. balyanboodja (Burgess et al. 2018), P. chesapeakensis (Man in ’t Veld et al. 2019), and P. personensis (Crous et al. 2020a) cluster within clade 6a of the Phytophthora phylogeny (Burgess et al. 2018). In a multigene phylogeny of the ITS, Btub, hsp90, coxI and nadh1 gene regions, P. aquae-cooljarloo differs from its sister taxon, P. gemini, by 8.8 %. Morphologically, P. aquae-cooljarloo is similar to other species in clade 6a, producing terminal, ellipsoid to ovoid, persistent, non-papillate sporangia, and it is also a high-temperature tolerant Phytophthora species. Isolates of P. aquae-cooljarloo are homothallic and produce abundant oospores in culture similarly to P. humicola, P. inundata, and P. condilina. Unlike P. balyanboodja, P. chesapeakensis and P. gemini, P. aquae-cooljarloo produces chlamydospores, but does not form any hyphal swellings which differs from all other species in the clade except P. balyanboodja. Phytophthora aquae-cooljarloo has been isolated over a 25-yr-period from seasonal ponds in the dry Banksia shrublands (the kwongan) in the sandplains north of Perth, Western Australia at a single location, Cooljarloo. Typus. AustrAliA, Western Australia, Cooljarloo, baited from pond water, collected by Department of Biosecurity, Conservation and Attractions, 20 Sept. 2017 (holotype MURU484, culture ex-type CBS 146550 = VHS36940, ITS, Btub, hsp90, coxI, nadh1 and LSU sequences GenBank MT210484, MT210475, MT210480, MT210466, MT210470 and MT210485, MycoBank MB835165). Additional materials examined. AustrAliA, Western Australia, Cooljarloo baited from pond water, collected by Department of Biosecurity, Conservation and Attractions, 18 Sept. 2019, cultures VHS39966, VHS39967; 1996, culture HSA2304. Colour illustrations. Typical kwongan vegetation, north of Perth, Western Australia (Photo: Giles Hardy). Typical ellipsoid and limoniform sporangia; aplerotic oogonia with paragynous antheridia; small chlamydospore; uniform colony on V8 agar. Scale bar = 20 μm. Bayesian inference tree based on a concatenated ITS, Btub, hsp90, coxI and nadh1 sequence alignment showing the placement of Phytophthora aquae-cooljarloo in Phytophthora Clade 6a. The tree was generated in MrBayes v. 3.2.6 (Huelsenbeck & Ronquist 2001) as a plugin in Geneious Prime® 2019.2.3 (www.geneious.com) using the GTR substitution model. The posterior probability values are shown at the nodes. The tree was rooted to P. thermophila (not shown) and the novel species is shown in bold font. Reza Mostowfizadeh-Ghalamfarsa, Department of Plant Protection, Shiraz University, Shiraz, Iran; e-mail: rmostowfi@shirazu.ac.ir Treena I. Burgess, Phytophthora Science and Management, Centre for Climate Impacted Terrestrial Ecosystems, Harry Butler Institute, Murdoch University, Murdoch, WA 6150, Australia; e-mail: tburgess@murdoch.edu.au © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 378 Persoonia – Volume 45, 2020 Phytopythium paucipapillatum 379 Fungal Planet description sheets Fungal Planet 1168 – 19 December 2020 Phytopythium paucipapillatum S.D. Langenhoven, W.J. Botha & L. Mostert, sp. nov. Etymology. The specific epithet refers to the sparsely papillated sporangia and oogonia. Classification — Pythiaceae, Pythiales, Oomycetes. Hyphae up to 5 µm thick, lacking hyphal swellings. Sporangia apical, unilaterally intercalary or perpendicular on sporangiophore, some sporangia clustered in groups of 3 – 5 at apex of sporangiophore, connected by short hyphal segments. Sporangia globose, subglobose, ovoid, obovoid, limoniform to ellipsoid or distorted shapes, 15–34 µm diam, most 19–25 µm diam. Sporangia mostly apapillate germinating directly, some papillate, internally proliferating with extended proliferation. Papilla apical or subapical, close to sporangiophore, 4 – 5 µm. Zoospores biflagellate, differentiated extrasporangially in an ephemeral vesicle, released through discharge tubes 3.7–5 µm wide, 7.5–11 µm long. Zoospore cysts spherical, 9–11 µm diam. Oogonia, small globose terminal, intercalary, some unilaterally intercalary, (18–)20–23(–26) (av. 22) µm diam, some oogonia ornamented with one to three short, blunt papillae. Antheridia up to three per oogonium, mostly monoclinous, or occasionally diclinous at a distance. Antheridia applied lengthwise to oogonium wall with a central fertilisation tube, antheridial cell 4 – 5 × 11– 20 µm with an undulating contour and one to several constrictions; some antheridia applied broadly apical to oogonium. Oospores plerotic or nearly so, (14 –)18 – 20(– 23) (av. 20) µm diam, wall thickness 0.9 –1.9 µm. Occasionally two oospores per oogonium. Ooplast 7–13 µm diam. Aplerotic index 76.9 %, ooplast index 54.4 %, oospore wall index 45.2 %. Cultural characteristics — Colony growth pattern on potato dextrose agar (PDA) and potato carrot agar (PCA) rosaceous, corn meal agar (CMA) slight aerial mycelium with coarsely radiate pattern and numerous micro tufts of aerial mycelium. Grows on PARP and PARPH selective media. Cardinal temperatures: min 10 °C, opt 25 °C, max 30 °C on PCA. Average growth rate at the optimum temperature was 8.55 mm/d for the STE-U isolates and 7.44 mm/d for MAFF 241149 on PCA. Growth study on CMA, min 10 °C, max 30 °C or between 30 °C and 35 °C for STE-U isolates and the MAFF isolate, respectively. The optimum growth temperature was 25 °C for STE-U 7843, 7844, 7847 and MAFF 241149. The optimum temperature for STE-U 7845, 7846 and 7848 was 30 °C. The average growth rate for the STE-U isolates with optimum growth temperatures at 25 °C and 30 °C were 9.77 mm/d and 10.63 mm/d, respectively. The average growth rate for the MAFF isolate was 8.79 mm/d. Typus. south AfricA, Western Cape Province, Wellington, Vitis sp. asymptomatic roots (Vitaceae), May 2013, S.D. Langenhoven (holotype and culture ex-type stored in a metabolically inactive state CBS 144082 = STE-U 7843; COI and ITS sequences GenBank KX372742 and KX372749, MycoBank MB819417). Additional materials examined. south AfricA, Western Cape Province, Wellington, grapevine roots (STE-U 7844, STE-U 7845, STE-U 7846, STE-U 7847, STE-U 7848). – JApAN, Nagano, uncultivated soil, collection date and collector unknown (as Ovatisporangium sp. 5, culture MAFF 241149). Notes — Phytopythium paucipapillatum sp. nov. was isolated from a nursery grapevine in South Africa. The inclusion of Ovatisporangium sp. 5 isolate MAFF 241149 in the species P. paucipapillatum is supported by morphological and phylogenetic data. Phylogenetically, P. paucipapillatum isolates formed a well-supported monophyletic clade with ITS (96 % bootstrap support and posterior probability of 1.0) and COI (99 % bootstrap support and posterior probability of 0.99). Phytopythium paucipapillatum was distinct from, but related to P. chamaehyphon, P. helicoides, P. fagopyri and Phytopythium sp. WJB-3 (of which only ITS is available). Morphological characteristics unique to P. paucipapillatum isolates were the plerotic and aplerotic oospores, compared to the mentioned closely related species with exclusively aplerotic oospores (Van der Plaats-Niterink 1981, McLeod et al. 2009, Baten et al. 2015). In addition, P. paucipapillatum is the only species with oogonial ornamentation as compared to P. chamaehyphon, P. helicoides, P. fagopyri and Phytopythium sp. WJB-3. Furthermore, the oogonia of P. paucipapillatum sometimes contain two oospores, unlike those of the above mentioned closely related species (including Phytopythium sp. WJB-3). Regarding internal proliferation, no nested proliferation was observed in P. paucipapillatum, as compared to P. chamaehyphon, P. fagopyri and P. helicoides – all of which display internal, nested proliferation. Furthermore, no internal proliferation has been observed for Phytopythium sp. WJB-3. Supplementary material Colour illustrations. Grapevine nursery, Wellington, South Africa. Colony growth on corn meal agar; sub-globose papillate sporangium; young, terminal multipapillate sporangium; zoospore discharge into a vesicle with a zoospore remaining in the sporangium; intercalary oogonium with monoclinous antheridium; oogonium with three antheridia attached; oogonia with papillation on its surface. Scale bar = 10 µm. FP1168-1 Maximum likelihood phylogeny of the internal transcribed spacernuclear ribosomal DNA region displaying species of the genus Phytopythium. Maximum likelihood analyses were performed in PhyML v. 3.3 (Guindon et al. 2010) under the best model (GTR+I+G for both ITS and COI ) as estimated using the Akaike Information Criterion in jModelTest v. 2 (Darriba et al. 2012). Support values were calculated from 100 bootstrap replicates. Maximum likelihood bootstrap percentages and Bayesian posterior probability values are indicated at the nodes. Support values less than 60 % bootstrap or 0.60 posterior probability are omitted or indicated with ‘–’. FP1168-2 Maximum likelihood phylogeny of the cytochrome c oxidase subunit 1 (COI) gene region displaying species of the genus Phytopythium. Both trees have been lodged in TreeBASE (study S22566). Shaun D. Langenhoven & Lizel Mostert, Department of Pant Pathology, University of Stellenbosch, Private Bag X1, Matieland, 7602, South Africa; e-mail: lmost@sun.ac.za & langenhovensd@gmail.com Chris F.J. Spies, ARC Plant Health and Protection, Private Bag X5017, Stellenbosch, 7599, South Africa; e-mail: SpiesC@arc.agric.za Wilhelm Botha, ARC Plant Health and Protection, Private Bag X134, Queenswood, Pretoria, 0121, South Africa; e-mail: BothaW@arc.agric.za Francois Halleen, Plant Protection Division, ARC Infruitec-Nietvoorbij, Private Bag X5026, Stellenbosch, 7599, South Africa; e-mail: HalleenF@arc.agric.za © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 380 Persoonia – Volume 45, 2020 Pseudopyricularia javanii 381 Fungal Planet description sheets Fungal Planet 1169 – 19 December 2020 Pseudopyricularia javanii A. Pordel & G. Ghorbani, sp. nov. Etymology. The species name is proposed in honour of Professor Mohammad Javan-Nikkhah, Iranian mycologist. Classification — Pyriculariaceae, Magnaporthales, Sordariomycetes. Mycelium on synthetic nutrient-poor agar (SNA), water agar (WA) supplemented with Cyperus leaves, and oatmeal agar (OA), consisting of smooth, hyaline, branched, septate hyphae. Conidiophores scattered, solitary, erect, pale brown, swollen at the base, macronematous, mononematous, typically unbranched, sometimes branched, straight, typically consisting of 0 – 5-septate, 35 – 85(–112) × 3 – 5 µm. Conidiogenous cells integrated, terminal, intercalary, sympodial, cylindrical, geniculate, denticulate; denticles cylindrical, thin-walled, pale brown. Conidia solitary, dry, obclavate, hyaline, (20 –)25 – 35(– 40) × 6 –7 µm, 2-septate, hilum often protuberant. Sexual morph unknown. Culture characteristics — Colonies on OA transparent, greenish olivaceous, reaching 34 mm diam after 1 wk at 23 – 25 °C; on potato dextrose agar (PDA) transparent, grey, and black reverse, reaching 37 mm diam after 1 wk at 23 – 25 °C. of new taxa), 2 978 characters including the alignment gaps were used. The phylogenetic tree suggested phylogenetic relatedness of the taxa from Iran to Pseudopyricularia with high statistical support (MLBP = 99 %). In the LSU sequences, the highest level of similarity (99.15 %; 819 /826) was to Ps. bothriochloae (reference sequence accession NG_058051.1), and Ps. hyrcaniana (99.27 %; 820 /826, GenBank KY457267), although the conidia in the new species is 2-septate. In species with 2-septate conidia, Ps. hagahagae has the highest level of similarity (98.87 %; 790 /799; reference sequence accession NG_059616), although the conidia of Ps. javanii are smaller than Ps. hagahagae (conidial size; (38 –)41– 45 (– 49) × (7–) 8 (– 9) µm). Pyricularia oryzae CBS 375.54 Pyricularia oryzae 100 Typus. irAN, Gilan Province, Someh Sara region, on infected leaves of Cyperus sp. (Cyperaceae), 15 Nov. 2018, A. Pordel (holotype in Iranian Research Institute of Plant Protection, IRAN 18060F, ex-type culture IRAN 3989C; ITS, LSU, CAL, RPB1 sequences GenBank MT472570, MT472574, MT472593 and MT472595, MycoBank MB837644). 100 CBS 365.52 Pyricularia oryzae CR0029 Pyricularia oryzae BF0028 Pyricularia oryzae JP0028 Pyricularia ctenantheicola GR0002 100 Pyricularia ctenantheicola GR0001 Pyricularia sp. CBS 133598 Pyricularia penniseticola 90 Pyricularia grisea 100 Additional material examined. irAN, Gilan Province, Someh Sara region, on infected leaves of Cyperus sp. (Cyperaceae), 15 Nov. 2018, A. Pordel (UTFC-PJ02; ITS, CAL, RPB1 sequences GenBank MT472569, MT472594 and MT472596). 100 Pyricularia grisea Pyricularia grisea BF0017 CR0024 JP0034 BR0029 Pyricularia zingibericola RN0001 Xenopyricularia zizaniicola Notes — Pseudopyricularia javanii is similar to Ps. higginsii, Ps. cyperi, Ps. iraniana, Ps. kyllingae, Ps. persiana, and Ps. hagahagae in having 2-septate conidia (Klaubauf et al. 2014, Pordel et al. 2017, Crous et al. 2018a). However, the conidia of Pseudopyricularia javanii are larger than those of Ps. higginsii, Ps. cyperi, Ps. kyllingae, and shorter than Ps. persiana, and Ps. hagahagae. It differs from Ps. iraniana in conidial shape, and size. To clarify the phylogeny of Ps. javanii within Pseudopyricularia, sequence data of CAL / ITS/RPB1 were combined. In the multi-gene analyses (gene boundaries of CAL: 1–723, ITS: 724 –1263, RPB1 1264 – 2265) of 39 isolates (37 taxa from NCBI and two sequenced specimens CBS 133593 100 Xenopyricularia zizaniicola Proxipyricularia zingiberis CBS 132356 CBS 132355 Neopyricularia commelinicola CBS 128303 CBS 128306 100 Neopyricularia commelinicola Pseudopyricularia higginsii 99 CBS 121934 Pseudopyricularia persiana UTFC-PO22 100 Pseudopyricularia persiana UTFC-PH01 90 Pseudopyricularia persiana UTFC-PO20 99 Pseudopyricularia kyllingae CBS 133597 100 Pseudopyricularia cyperi Pseudopyricularia cyperi CBS 665. 79 CBS 133595 Pseudopyricularia iraniana 100 UTFC-PO15 Pseudopyricularia iraniana Pseudopyricularia iraniana 93 100 Pseudopyricularia javanii 99 Pseudopyricularia javanii 100 97 UTFC-PO14 UTFC-PO13 UTFC-PJ02 IRAN 3989C Pseudopyricularia hedjaroudii UTFC-PH02 Pseudopyricularia hedjaroudii UTFC-PH01 98 Pseudopyricularia hyrcaniana UTFC-PO10 100 Pseudopyricularia hyrcaniana UTFC-PO12 98 Pseudopyricularia hyrcaniana UTFC-PO11 Macgarvieomyces borealis 95 Macgarvieomyces juncicola Bambusicularia brunnea CBS 461.65 CBS 610.82 CBS 133599 Barretomyces calatheae CBS 129274 0.02 Colour illustrations. Cyperus growing in Iran. Solitary, erect, unbranched conidiophore; obclavate conidia. Scale bars = 10 µm. Maximum Likelihood tree inferred with MEGA v. 6 software (Tamura et al. 2013) from the combined CAL, ITS and RPB1 gene regions of 39 isolates. The novel species is shown in bold. Bootstrap support values from ML ≥ 90 % are provided above internodes. Adel Pordel, Plant Protection Research Department, Baluchestan Agricultural and Natural Resources Research and Education Centre, AREEO, Iranshahr, Iran; e-mail: a_pordel@areeo.ac.ir Golzar Ghorbani, Department of Plant Protection, College of Agriculture and Natural Resources, University of Tehran, Karaj 31587-77871, Iran; e-mail: Golzar.ghorbani@ut.ac.ir © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 382 Persoonia – Volume 45, 2020 Pseudosubramaniomyces septatus 383 Fungal Planet description sheets Fungal Planet 1170 – 19 December 2020 Pseudosubramaniomyces septatus Torres-Garcia, Gené, Dania García, sp. nov. Kirk 1982). The genus is characterised by having solitary, unbranched conidiophores and terminal, polyblastic, denticulate conidiogenous cells, which give rise to catenate conidia. It resembles Subramaniomyces but differs by the lack of lateral conidiogenous cells and tends to have pale brown conidiophores, in contrast to the dark brown conidiophores observed in Subramaniomyces (Varghese & Rao 1980, Crous et al. 2017a). Of note however is that the presence of dimorphic conidia (i.e., hyaline to pale brown, ellipsoidal to broadly fusoid intercalary conidia vs elongate fusoid to acicular and brown terminal conidia), typical of S. fusisaprophyticus (Kirk 1982) and also observed in P. septatus, was not mentioned in the protologue of Pseudosubramaniomyces. Pseudosubramaniomyces septatus differs from P. fusisaprophyticus and other accepted species of Subramaniomyces (Varghese & Rao 1980, Braun & Hill 2002, Da Cruz et al. 2007, Crous et al. 2017a) mainly by the presence 0 –2-septate intercalary conidia. Furthermore, P. septatus has longer intercalary (12 – 24.5 µm) and terminal (27– 34 µm) conidia than those of P. fusisaprophyticus, which measure (13 –)17–18.5(– 21) µm and (18 –)25 – 31 µm long, respectively (Kirk 1982). Etymology. Name refers to the presence of septate conidia. Classification — Beltraniaceae, Xylariales, Sordariomycetes. On potato carrot agar (PCA) at 25 °C. Mycelium partly superficial, partly immersed, composed of branched, septate, pale brown, smooth hyphae, 1.5–2 µm wide. Conidiophores solitary, erect, unbranched, septate, pale brown at the base, hyaline at the apex, smooth, subcylindrical, 9 – 66 × 2 – 3 µm. Conidiogenous cells integrated, terminal, polyblastic, sympodial, denticulate, with up to five denticles, hyaline, smooth, subcylindrical, 14 – 22 × 2 – 2.5 µm. Conidia dry, at first solitary, latter forming short branched or unbranched chains, dimorphic; apical conidia aseptate, pale brown, smooth, cylindrical to subcylindrical, with obtuse apex and truncate base, 27–34 × 2.5–4 µm; intercalary conidia (including ramoconidia), 0 –1(– 2)-septate, hyaline to subhyaline, smooth, fusoid or navicular, 12 – 24.5 × 2 – 4 µm. Sexual morph not observed. Culture characteristics at 25 °C after 1 wk — Colonies on PCA reaching 12 –16 mm diam, slightly elevated, dull green (30E4) to white (1A1) (Kornerup & Wanscher 1978), velvety, regular margin; reverse dull green (30E4) to white (1A1); sporulation sparse. On potato dextrose agar (PDA) reaching 21 mm diam, slightly elevated, greyish brown (5E3) to white (1A1), velvety, regular margin; reverse yellowish white (3A2); sporulation absent. On oatmeal agar (OA) reaching 9 –12 mm diam, flat, white (1A1), velvety, regular margin; reverse brownish grey (4F2) to greyish yellow (4C5); sporulation absent. Cardinal temperature for growth — Opt 25 °C, max 30 °C, min 5 °C. Our phylogenetic analysis using the barcodes LSU and ITS places P. septatus close to the species P. fusisaprophyticus in the family Beltraniaceae. A megablast search using LSU sequences shows that P. septatus has a similarity of 98.27 % (737/750) with P. fusisaprophyticus (CBS 418.95; GenBank EU040241.1) and 97.60 % (732/750) with Beltraniopsis neolitseae (CBS 137974; GenBank MH878610.1); meanwhile the similarity using ITS barcode was 91.24 % (500/548) with P. fusisaprophyticus (CBS 418.95; GenBank EU040241.1) and 90.42 % (500/553) with B. neolitseae (CBS 137974; GenBank NR148072.1). Typus. spAiN, Catalonia, Barcelona province, Montseny Natural Park, El Sot de l’Infern stream, fluvial sediments, Oct. 2018, D. Torres-Garcia (holotype FMR H-17583, culture ex-type FMR 17583, also in CBS; LSU and ITS sequences GenBank LR700217 and LR700216, MycoBank MB837574). Notes — Pseudosubramaniomyces was proposed by Crous et al. (2017a) based on Subramaniomyces fusisaprophyticus (= Ramularia fusisaprophytica), a fungus usually isolated from decaying leaves of different trees (Matsushima 1971, B. neolitseae CBS 137974T 100 B. longiconidiophora MFLUCC 17-2139T Beltraniopsis B. pseudoportoricensis CBS 145547T 75 B. humicola CBS 203.64 76 B. carolinensis 9502 IFO B. endiandrae CPC 22193 71 85 98 Beltraniella B. rhombica CBS 123.58 B. krabiensis MFLUCC 16-0257T Beltrania B. pseudorhombica CBS 138003T 100 100 P. fusisaprophyticus CBS 418.95T P. septatus FMR 17583T 100 Beltraniaceae B. thailandica MFLUCC 16-0377 100 P. inaequiseptata MUCL 41089 Pseudosubramaniomyces Parapleurotheciopsis Subramaniomyces podocarpi CBS 143176T 100 Seimatosporium elegans NBRC 32674 Seimatosporium eucalypti CPC 159 0.02 Colour illustrations. Montseny Natural Park, Catalonia, Spain. Colony on PDA and PCA after 7 d at 25 °C; conidiophores and conidia after 14 d at 25 °C. Scale bars = 25 µm (habitat in PCA), 10 µm (microscopic structures in PCA). Phylogenetic tree based on Maximum likelihood analysis obtained by RAxML using the combined LSU and ITS sequences of Pseudosubramaniomyces and related genera in the family Beltraniaceae. Bootstrap support values above 70 % are indicated on the nodes. The alignment included 1 530 bp and was performed using Kimura-2 parameter Gamma distribution with Invariant sites (G+I) as the best nucleotide substitution model. The tree was rooted with Seimatosporium elegans NBRC 32674 and Seimatosporium eucalypti CPC 159. The alignment was constructed with MEGA v. 6 software (Tamura et al. 2013). The new species proposed in this study is indicated in bold face. A superscript T denotes ex-type cultures. Daniel Torres-Garcia, Josepa Gené & Dania García, Mycology Unit, Medical School and IISPV, Universitat Rovira i Virgili, Sant Llorenç 21, 43201 Reus, Tarragona, Spain; e-mail: daniel.torres@urv.cat, josepa.gene@urv.cat & dania.garcias@urv.cat © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 384 Persoonia – Volume 45, 2020 Pyrenochaetopsis rajhradensis 385 Fungal Planet description sheets Fungal Planet 1171 – 19 December 2020 Pyrenochaetopsis rajhradensis Spetik, Eichmeier & Berraf-Tebbal, sp. nov. MT453283.1; Identities = 471/471 (100 %), no gaps), Phoma sp. (GenBank MN401018.1; Identities = 471/471 (100 %), no gaps) and Pyrenochaetopsis leptospora (GenBank LR216648.1; Identities = 471/471 (100 %), no gaps). The closest hits using the LSU sequence had the highest similarity to Pyrenochaetopsis microspora (GenBank NG_069864.1; Identities = 937/939 (99 %), no gaps), Pyrenochaetopsis leptospora (GenBank NG_069858.1; Identities = 937/939 (99 %), no gaps) and Pyrenochaetopsis sp. (GenBank KJ395496.1; Identities = 937/939 (99 %), no gaps); closest hits using the rpb2 sequence are Pyrenochaetopsis leptospora (GenBank LT623283.1; Identities = 893/906 (99 %), no gaps and GenBank LT623282.1; Identities = 846/858 (99 %), no gaps), Phaeopoacea festucae (GenBank MF795835.1; Identities = 840/854 (98 %), 2/854 (0 %)) and Pyrenochaetopsis poae (GenBank LT623286.1; Identities = 864/906 (95 %), no gaps). The closest hits using the tef1-α sequence had the highest similarity to Pyrenochaetopsis leptospora (GenBank MF795881.1; Identities = 266/281 (95 %), 3/281 gaps (1 %)), Parastagonospora novozelandica (GenBank MK540151.1; Identities = 57/58 (98 %), no gaps) and Parafenestella rosacearum (GenBank MK357586.1; Identities = 186/245 (76 %), 27/245 gaps (11 %)). The closest hits using the tub2 sequence had the highest similarity to Pyrenochaetopsis poae (GenBank KJ869243.1; Identities = 407/407 (100 %), no gaps), Pyrenochaetopsis leptospora (GenBank MF795917.1; Identities = 402/407 (99 %), no gaps and GenBank LT623242.1; Identities = 325/332 (98 %), no gaps). Etymology. Named after Rajhrad (Czech Republic) where the fungus was collected. Classification — Pyrenochaetopsidaceae, Pleosporales, Dothideomycetes. Conidiomata pycnidial, brown, solitary or aggregated, semiimmersed, globose to ovoid, setose, ostiolate, uniloculate. Conidiogenous cells phialidic, hyaline, discrete and integrated. Conidia hyaline, aseptate, cylindrical to allantoid, guttulate, (3.6 –)4.1– 4.9(– 5.7) × (1.4 –)1.6 – 2.2(– 2.4) µm (av. ± S.D. 4.5 ± 0.4 × 1.8 ± 0.2 µm, L/W ratio = 2.5). Sexual morph unknown. Culture characteristics — Colonies on potato dextrose agar (PDA) reaching 23.8 mm diam at 25 °C after 10 d, margin regular, floccose, dirty white; reverse white. On malt extract agar (MEA) reaching 22 mm diam after 10 d, margin regular, floccose, dirty white; reverse white. On oatmeal agar (OA) reaching 25.8 mm diam after 10 d, margin regular, floccose, white; reverse white. Typus. czech republic, Rajhrad, isolated as saprobe from dead wood of Buxus sempervirens (Buxaceae), July 2018, M. Spetik (holotype CBS H-24478, ex-type culture CBS 146846 = MEND-F-51, ITS, LSU, rpb2, tef1 and tub2 sequences GenBank MT853115, MT853182, MT857727, MT857725 and MT857726, MycoBank MB836856). Notes — Based on a megablast search of NCBI nucleotide database, the closest hits using the ITS sequence had the highest similarity to Pyrenochaetopsis leptospora (GenBank 70/81 98/96 56/61 CBS 101635 95/97 -/51 Pyrenochaetopsis rajhradensis sp. nov. CBS 146846 = MEND-F-51 CBS 136769 Pyr. poae CBS 122787 Pyrenochaetopsis leptospora CBS 119739 Pyr. setosissima -/52 FMR 13688 Pyr. microspora FMR 13715 Pyr. americana 59/66 CBS 143034 Pyr. globosa FMR 13769 Pyr. uberiformis 69/79 87/ 90 FMR 13683 Pyr. paucisetosa 99/100 93/100 98/100 66/92 CBS 139506 Pyr. tabarestanensis CBS 142458 FMR 13781 Pyr. botulispora FMR 13790 CBS 142459 Pyr. confluens CBS 343 85 Pyr. decipiens 92/86 67/68 CBS 12445 Pyr. indica CBS 44581 Xenopyrenochaetopsis pratorum CBS 143033 Neopyrenochaetopsis hominis 0.02 Maximum likelihood tree obtained from the ITS, tub2, LSU and rpb2 gene sequences of Pyrenochaetopsis species of our isolates and sequences retrieved from GenBank. The tree was built using MEGA v. 7.0 (Kumar et al. 2016). The combined LSU, ITS, tub2 and rpb2 sequence data set consisted of 17 Pyrenochaetopsis strains with Xenopyrenochaetopsis pratorum and Neopyrenochaetopsis hominis as the outgroup taxa and consisted of 2 195 characters. Of these 1 643 were constant, 188 were variable and parsimonyuninformative and 337 were parsimony-informative. A heuristic search of these 337 parsimony-informative characters resulted in 1 000 equally parsimonious trees of 467 steps with CI = 0.72, RI = 0.65 and HI = 0.28. The ML analysis yielded a best scoring tree with the final ML optimization likelihood value of –4554.41 (ln) and a gamma distribution shape parameter value of α = 0.1411. All individual trees obtained from single gene datasets were essentially similar in topology and not substantially different from the tree generated from the concatenated dataset. One of the two ML trees obtained is presented with ML/MP bootstrap support values at the nodes. The alignment and tree are available in TreeBASE (Submission ID: 26835). Colour illustrations. Buxus sempervirens growing in Lednice castle garden. Pycnidia forming on sterile poplar twig on WA; pycnidium in culture oozing conidia; conidiogenous cells; conidia. Scale bars = 200 µm (pycnidia), 10 µm (all others). Milan Spetik, Akila Berraf-Tebbal & Ales Eichmeier, MENDELEUM – Institute of Genetics, Mendel University in Brno, Valticka 334, 69144, Czech Republic; e-mail: milan.spetik@mendelu.cz, ales.eichmeier@mendelu.cz & qqberraf@mendelu.cz Alla Eddine Mahamedi, Laboratoire de Biologie des Systèmes Microbiens (LBSM), Département des Sciences Naturelles, Ecole Normale Supérieure de Kouba, Alger BP 92, Vieux-Kouba, Alger, Algeria; e-mail: aladin1342@yahoo.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 386 Persoonia – Volume 45, 2020 Russula shawarensis 387 Fungal Planet description sheets Fungal Planet 1172 – 19 December 2020 Russula shawarensis Kiran & Khalid, sp. nov. Etymology. The specific epithet, shawarensis, refers to the Shawar Valley, the locality from where the type was collected. Classification — Russulaceae, Russulales, Agaricomycetes. Pileus medium to large-sized, 40 – 90 mm diam, semi-globose, convex to hemispheric, expanding plane and centrally slightly depressed; cuticle thin, adnate, hardly peeling, areolate; margin incurved and entire when young, later recurved, when mature radially splitting; surface matt, smooth, light pinkish brown (9.3 R 6.4/0.9) to grey buff (4.1 GY 7/0.3), discolouring to light brown (0.6Y 5.8/2.2) (Kornerup & Wanscher 1978). Lamellae moderately distant, adnate-emarginate, equal, frequently forking, brittle, cream white, spotted light yellow-brown after handling, lamellulae very rare to absent, edge even, concolorous. Stipe 25 – 60 × 10 –12 mm, obclavate, central, velvety, white with yellow-brown to brown spots, especially near the base. Context white, changing to yellowish brown upon bruising, compact. Spores (5.9 –)6.6 –7.8(– 9.2) × (5.1–)5.7– 6.7(– 8.6) μm, av. 7.2 × 6.2 μm, subglobose to broadly ellipsoid, Q = (1.0 –)1.08 –1.26(–1.4), Qav = 1.17; ornamentation of small, distant to moderately distant (4 – 5(– 6) in a 3 μm diam circle) amyloid spines, (0.8 –)0.9 –1.1(–1.2) μm high, radially oriented from suprahilar spot, occasionally fused in pairs, ((0–)1(–2) fusions in the circle), connected by dispersed fine line connections ((0–)1–2(–3) in the circle); suprahilar spot distinct, not amyloid or with few small amyloid dots. Basidia (31.5 –)34 – 39.5 (– 41) × (7–)9 –10.5(–11) μm, av. 37 × 10 μm, 2 – 4-spored, clavate; basidiola first cylindrical or ellipsoid, then clavate, c. 4 –10 μm wide. Hymenial cystidia on lamellar sides widely dispersed, 200–300/mm2, (66–)72–92.5(–105) × (9.5–)10–13(–14) μm, av. 82.1 × 11.4 μm, fusiform or rarely clavate, often pedicellate, apically acute or sometimes obtuse and mostly with 4 –17 μm long appendage; contents heteromorphous, crystalline-banded, turning brown to almost greyish black in sulfovanillin; abundant near the lamellae edges, (55 –)61–76(– 88) × (6 –)8 –11.5(– 13.5) μm, av. 68.5 × 9.6 μm, more frequently clavate, sometimes also cylindrical, usually obtuse, frequently apically constricted or appendiculate. Lamellar edges fertile; marginal cells not well differentiated, smaller, c. 10 – 20 × 4 – 5 μm, cylindrical or clavate. Pileipellis orthochromatic in Cresyl blue, not sharply delimited from the underlying context, 175 – 200 μm deep, strongly gelatinised; suprapellis 45 – 55 μm deep, disconnected, of dense, ascending and near the surface repent hyphal terminations; gradually passing to 45–120 μm deep subpellis of irregularly oriented, intricate, (2.5–)3–4(–4.5) μm wide hyphae. Acid-resistant incrustations absent. Hyphal terminations in pilColour illustrations. Quercus floribunda dominated forest in Lower Shawar (Khyber Pakhtunkhwa province, Pakistan) where the holotype was collected. Left top: Pileal surface of collection LAH 35453. Centre bottom: basidiomata of collection LAH 35452. Right top: Scanning electron photograph of spores from LAH 35452. Line drawings all from the holotype LAH 35453. Right bottom: basidia and basidiola (left top), marginal cells (centre) and spores (left bottom), hymenial cystidia near the lamellae edges (right top) and lamellae sides (right bottom). Left bottom: pileocystidia near the pileus centre (left top) and near the pileus margin (left bottom), hyphal terminations near the pileus centre (right top) and near the pileus margin (right bottom) from holotype. Scale bars = 10 mm (basidiomata), 5 μm (spores), 10 μm (all other microscopic structures). eipellis near the pileus margin, composed of 2 – 3 unbranched cells with the basal cell often shorter and inflated, often slightly flexuous, thin-walled; terminal cells (23 –)33 – 63.5(– 81) × (2.5 –)3 – 4.5(– 5) μm, av. 48.2 × 3.8 μm, mainly cylindrical, apically often slightly attenuated; subterminal cells usually equally wide and sometimes shorter, usually unbranched. Hyphal terminations near the pileus centre slightly smaller, terminal cells (23–)32–51(–64.5) × (3–)3.5–5(–7) μm, av. 41.5 × 4.2 μm, subterminal and lower cells more frequently inflated. Pileocystidia near the pileus margin very abundant, mainly onecelled, cylindrical to narrowly clavate, thin-walled, terminal cells (31.5 –)33 – 90(–155) × (4.5 –)6 – 8(– 9) μm, av. 61.6 × 6.7 μm, apically mainly obtuse, contents heteromorphous, slowly turning greyish in sulfovanillin. Pileocystidia near the pileus centre similar, terminal cells (22 –) 30 – 86(–166) × (5 –)6 – 8.5(– 9.5) μm, av. 58 × 7 μm. Cystidioid hyphae in subpellis and context dispersed, contents heteromorphous-banded. Typus. pAkistAN, Khyber Pakhtunkhwa province, Malakand division, Swat district, Lower Shawar, alt. 1 200 m, on the floor of Quercus floribunda dominated moist temperate forest mixed with a few pines, 7 Sept. 2015, Z. Ullah & M. Kiran MK-KS49 (holotype LAH 35453, ITS and LSU sequences GenBank MT738294 and MT738269, MycoBank MB836118). Additional materials examined. pAkistAN, Khyber Pakhtunkhwa province, Malakand division, Swat district, Lower Shawar, alt. 1 200 m, on the floor of Quercus floribunda dominated moist temperate forest mixed with a few pines, 7 Sept. 2015, Z. Ullah MK-KS26 (LAH 35452, ITS, LSU and rpb2 sequences GenBank MT738291, MT738266 and MT732175,); ibid., 26 July 2018, Z. Ullah & J. Khan AS 48 (LAH36424, ITS, LSU and rpb2 sequences GenBank MT738290, MT738265 and MT732174); AS 61 (LAH36425, ITS, LSU and rpb2 sequences GenBank MT738292, MT738267 and MT732176); AS 75 (LAH36426, ITS, LSU and rpb2 sequences GenBank MT738293, MT738268 and MT732177). Notes — The ITS sequence of the type collection has the closest GenBank BLAST match (97.8 %) with a sequence identified as Russula atroglauca originating from Kyrgyzstan (GenBank MK351735). More than 80 sequences that fall in the UNITE species hypothesis SH1423803.08FU of R. atroglauca (www/unite.ut.ee) are within 96 % GenBank BLAST identity. To distinguish the Pakistani collections from R. atroglauca and other European species of the section Griseinae, we performed phylogenetic multilocus analysis of ITS, LSU and rpb2 regions (for the phylogenetic tree and analysed sequences see Supplementary material FP1172-1+2). To avoid misidentification (text continues on Supplementary material page FP1172) Supplementary material FP1172-1 Maximum likelihood phylogeny estimated for members of subsection Griseinae inferred from ITS, LSU and rpb2 regions in RAxML-173 HPC2 v. 8.2.10 (Stamatakis 2015), with rapid bootstrapping (1 000 iterations), computed on the CIPRES web server (www.phylo.org; Miller et al. 2010) under default settings including a General Time Reversible (GTR) + Gamma (G) model of sequence evolution. Bootstrap support values followed by Bayesian posterior probabilities computed in MrBayes v. 3.2 (Ronquist et al. 2012) are indicated at the nodes. Species names are followed by herbarium codes and country of origin. FP1172-2 List of samples and sequences used. Munazza Kiran, Zia Ullah & Abdul Nasir Khalid, Department of Botany, University of the Punjab, Quaid-e-Azam campus, Lahore 5090, Pakistan; e-mail: munazzakiran@gmail.com, ziaullah.phd.mmg@pu.edu.pk & drankhalid@gmail.com Miroslav Caboň, Institute of Botany, Plant Science and Biodiversity Centre, Slovak Academy of Sciences, Dúbravská cesta 9, SK-84523, Bratislava, Slovakia; e-mail: miroslav.cabon@gmail.com Felix Hampe, Wetzlarer Strasse 1, 35510 Butzbach, Germany; e-mail: felix.hampe@email.de © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 388 Persoonia – Volume 45, 2020 Serendipita whamiae 389 Fungal Planet description sheets Fungal Planet 1173 – 19 December 2020 Serendipita whamiae Dearnaley, T.W. May & Linde, sp. nov. Notes — Serendipita is a genus of Agaricomycetous fungi, many of which occur as endophytes in the roots of grasses, ericoids, liverworts and orchids (Weiss et al. 2016). The group is characterised by longitudinally septate basidia, long, wormlike basidiospores and DNA sequence data (Oberwinkler 1964, Roberts 1993, Basiewicz et al. 2012, Riess et al. 2014). Serendipita whamiae is a new species of Serendipitaceae with morphological similarities to the type species, S. vermifera, including probasidia 5 – 9 µm, longitudinally septate basidia and vermiform basidiospores, although the latter are shorter (11– 50 µm) than that described by Oberwinkler (1964), Warcup & Talbot (1967) and Roberts (1993) at 30 – 60 µm, 45 – 64 µm, 21– 86 µm, respectively. Compared to sequenced, named species within Serendipita, S. whamiae is distinct on BLAST matches from S. herbamans (ITS; 87 % identity over 637 bp; GenBank NR_144842) and S. indica (ITS; 68 % identity over 685 bp; GenBank NR_166023) and also from the type of the genus, S. vermifera (LSU; 87 % identity over 568 bp; GenBank HM030724). Otherwise, there is a series of unnamed environmental and endophyte sequences (EU625995 to KY798218) that form a well-supported clade including the sequence from the type of S. whamiae (all within 96 % similarity) that could prove conspecific, but for which morphological data is not available. Whitehead et al. (2017) found that Serendipita defined by multigene species delimitation had maximum within species variation of 4.1 % for ITS. Etymology. Named in honour of the well-known naturalist of the Stanthorpe region, Dell Wham. Classification — Serendipitaceae, Sebacinales, Agaricomycetes. Sporophore produced by the soil on agar method (Warcup & Talbot 1967), grey, resupinate hyphae occurring loosely on the surface of soil clods. Probasidia globose 5–9 µm to subglobose 7–9 × 6–8 µm diam, some with sub-basidial cells. Metabasidia crucially septate, in groups of 2–3 on short stalks from hyphae, globose, 7 µm diam, to subglobose, 6 – 8 µm diam. Basidia ovate, 7–11 × 6 –7 µm diam, longitudinally septate, with 2 – 4 sterigmata. Sterigmata 5–21 µm long, narrowing at apex. Basidiospores vermiform, some with septa, 11– 50 × 1– 2 µm diam. Culture characteristics — Colonies on potato dextrose agar (PDA) up to 6 cm diam after 3 wk growth at 22 °C, pinkish buff, flattened, without aerial mycelium, margins irregular, surface wrinkled in the central part, reverse pinkish buff. Hyphae hyaline, thin-walled, lacking clamps, 2 µm in width. Monilioid cells globose, 7 µm diam to subglobose, 6 –11 × 5 –10 µm diam, in chains. Typus. AustrAliA, Queensland, Stanthorpe, Girraween National Park, open Eucalyptus woodland, S27°49'13" E151°58'47", alt. 1008 m, isolated as an endophyte from roots of Eriochilus cucullatus (Orchidaceae), 7 Apr. 2016, J.D.W. Dearnaley EC3A (holotype BRIP 71159 living culture stored in a metabolically inactive state, ITS and LSU sequences GenBank KY798218 and MT422063, MycoBank MB835492). KF000449 Helvellosebacina sp. JQ665545 Sebacina incrustans KY524361 OTUF from Elythranthera brunonis 0.91 1 KF061288 OTUD from Caladenia tentaculata KY524370 OTUE from Caladenia flava 0.89 0.99 KY524368 OTUH from Caladenia cairnsiana 0.79 KY524420 OTUG from Caladenia procera 1 MN872348 OTUB from Glossodia major FJ611950 OTUA from Caladenia atroclavia 1 0.81 0.79 1 0.99 EU625991 from Eriochilus cucullatus EU625992 from Eriochilus cucullatus MT423705 from Bulbophyllum globuliforme MT423704 from Eriochilus autumnalis NR_166023 Serendipita indica Bayesian inference tree of ITS sequences from S. whamiae (bold) and related Sebacinales species in GenBank using MrBayes v. 3.2.7 (Huelsenbeck & Ronquist 2001) as implemented in Geneious v. 10.2.6 (Kearse et al. 2012). ClustalW was used for the alignment. The numbers above the branches are Bayesian posterior probabilities with values less than 0.7 not shown. Helvellosebacina sp. (KF000449) and Sebacina incrustans (JQ66545) were used as outgroups. Sequences labelled OTUA etc. are Australian orchid isolates categorised by Whitehead et al. (2017). Sequences in red are named species of Serendipita. 0.94 NR_144842 Serendipita herbamans FJ788844 from Pterygodium catholicum 1 0.91 EU625995 from Eriochilus scaber FN663149 from Eriochilus scaber 0.98 0.72 KF061290 from Phyllanthus calycinus EU626000 from Cyrtostylis reniformis MH633971 from Festuca rubra 0.91 0.93 HQ154261 from Trifolium repens KJ188479 from Neottia ovata KY798218 Serendipita whamiae KF061298 from Microtis unifolia EU625989 from Cavendishia nobilis Colour illustrations. Eriochilus cucullatus (inset) in Eucalyptus woodland at Girraween National Park (Photo credit Ian Milinovich). Serendipita whamiae (clockwise from top left) colony on PDA; monilioid cells; probasidia; metabasidia; basidium; basidiospores. Scale bars = 1 cm (colony and inset), 10 μm (all others). 0.09 John D.W. Dearnaley, Centre for Crop Health, University of Southern Queensland, Toowoomba 4350, Queensland, Australia; e-mail: john.dearnaley@usq.edu.au Tom W. May, Royal Botanic Gardens Victoria, Birdwood Ave, Melbourne, VIC 3004, Australia; e-mail: tom.may@rbg.vic.gov.au Celeste Linde, Ecology and Evolution, Research School of Biology, College of Science, The Australian National University, Canberra, ACT, 2601, Australia; e-mail: celeste.linde@anu.edu.au © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 390 Persoonia – Volume 45, 2020 Suhomyces rilaensis 391 Fungal Planet description sheets Fungal Planet 1174 – 19 December 2020 Suhomyces rilaensis R.A. Dimitrov & Gouliamova, sp. nov. Etymology. Ri-la-en-sis, referring to the locality Rila National Park from which this species was is olated. Classification — Incertae sedis, Saccharomycetales, Saccharomycetes. After 7 d at 25 °C in 5 % glucose-yeast extract broth, cells are spherical, subglobose, ellipsoidal and oblong, 2 –7 × 2–9 μm, occurring singly or in clusters. Asexual reproduction is by multilateral budding. After 7 d at 25 °C on 5 % malt extract agar (MEA) the culture is cream, butyrous, smooth, glistening, convex and with an entire margin fringed with filaments. Dalmau plate culture after 10 d on yeast morphology agar results in the formation of pseudohyphae. Aerobic growth is dimorphic, center is erased, and the margin is completely eroded, fringed with filaments. Ascospore production was not detected either alone or in pairs on yeast extract, malt extract agar (YMA), 5 % MEA, McClary acetate agar, potato dextrose agar (PDA), malt agar (MA2) and diluted V8 agar. Fermentation — Glucose is fermented. Galactose, maltose, sucrose, lactose and raffinose are not fermented. Carbon assimilation — D-glucose, D-galactose, D-glucosamine (+,w), D-ribose, D-xylose, α,α-trehalose, cellobiose (+,w), salicin, arbutin, glycerol, meso-erythritol, ribitol, xylitol (+,w), D-glucitol, D-mannitol, Glucono b-lactone (+,w), 2 keto-D-gluconate, D-gluconate (w,-), succinate, citrate, ethanol and propane 1,2 diol are assimilated. L-Sorbose, L-arabinose, D-arabinose, L-rhamnose, sucrose, maltose, methyl α –glucoside, melibiose, lactose, raffinose, melezitose, inuline, soluble starch, galactitol, myo-inositol, D-glucuronate, D-galactouronate, DL-lactate, methanol, butane 2,3 diol, quinic acid, saccharate and galactonic acid are not assimilated. Nitrogen assimilation — Nitrite, ethylamine, L-lysine are assimilated. Nitrate, creatine, creatinine, N-acetyl- glucosamine and imidazole are not assimilated. Other tests — Growth in medium containing 0.01 and 0.1 % cyclohexemide is negative. Growth in medium containing 50 % and 60 % glucose is negative. Starch production, urea and DBB tests are negative. Growth in medium containing 10 % NaCl is positive. Growth in 15 % NaCl is negative. Growth at 25 °C, 30 °C, 35 °C and 37 °C is positive. Growth at 42 °C is negative. Growth without all vitamins test is negative. Notes — Suhomyces tanzawaensis was isolated from mosses in Japan (Nakase et al. 1988) and had no known close relatives for a long time. In 2001 six additional species were isolated from mushrooms, plants and insect frass (Kurtzman 2001). Suh et al. (2004) isolated 16 new yeast species belonging to the clade from the gut of mushroom feeding insects (Suh et al. 2004). Kurtzman et al. (2016) proposed a new genus Suhomyces to accommodate members of the clade. During a yeast biodiversity survey conducted in Bulgaria in 2008 – 2011 three conspecific yeast strains (100 % identity in both LSU nrDNA and ITS nrDNA sequences) belonging to the genus Suhomyces were isolated from the gut of beetles. Three strains, DZ3, D301 and DKUZ, have the most similar sequences in the database belonging to S. bolitotheri (97 % identity in LSU nrDNA sequence) and S. tanzawaensis (88 % identity in ITS1+2 nrDNA sequence), thus indicating that the three Bulgarian strains represent a new yeast species. Phylogenetic analysis of combined LSU rDNA and ITS sequences placed the new species and S. bolitotheri in a separate subclade (100 % support). Pairwise analysis of the sequences from multiple alignment data showed that the new strains show 73 % similarity (242 identical nt.: 626 subst., 36 gaps) in ITS-LSU nrDNA with S. bolitotheri and 75 % similarity (248 identical nt.: 622 subst., 86 gaps) with S. choktaworum. The results of the phylogenetic analyses were confirmed by the comparative analysis of physiological profiles of the yeast strains and closest relatives on the phylogenetic tree. The analysis showed that six physiological characteristics distinguish the new strains from S. bolithoteri. The new species is not able to ferment galactose, is able to assimilate propane 1,2 diol and is not able to assimilate L-sorbose. Growth in the presence of 16 % NaCl, 50 % and 60 % glucose, and in the presence of 0.01 cycloheximide is negative. Nine characteristics distinguished the new strains from S. choctaworum. The new species is not able to ferment galactose, is not able to assimilate L-sorbose, L-arabinose and D-arabinose. Growth in the presence of 16 % NaCl, 50 % and 60 % glucose, and in the presence of 0.01 and 0.1 % of cycloheximide is negative. Typus. bulGAriA, in Podgorie area below Samuilovo village, from the gut of Bolitophagus interruptus found on a Polyporus sp., D. Gouliamova (holotype DZ3 preserved in metabolically inactive state in the yeast collection of the Institute of Microbiology, Sofia, Bulgaria. The ex-type culture is deposited at National bank for microorganisms and cell cultures (NBIMCC), Sofia Bulgaria, and at the CBS-KNAW culture collection of Westerdijk Fungal Biodiversity Institute, Utrecht, the Netherlands as NBIMCC 8930 = CBS 12453; D1/D2 LSU and ITS sequences GenBank HM627113 and HM627148, MycoBank MB802451). Additional materials examined. bulGAriA, in vicinity of Rila monastery, strain isolated from the gut of Bolitophagus reticulatus, D301 = NBIMCC 8929 = CBS 12443, D1/D2 LSU and ITS sequences GenBank HM627061 and HM627147; DKUZ1 isolated from unidentified grasshopper (Orthoptera), NBIMCC 8931 = CBS 12460, D1/D2 LSU and ITS sequences GenBank HM627116 and HM627150. Colour illustrations. A view of reservoir Koprinka in Rose Valey, Bulgaria. Morphology of cells of Suhomyces rilaensis DZ3T in 5 % glucose broth after 1 wk; Bolitophagus interruptus (Photo credits to S. Zayakov and K. Makarov, https://www.zin.ru/Animalia/Coleoptera/eng/bolintkm.htm). Scale bar = 5 μm. Phylogenetic tree obtained by the analysis of combined ITS and LSU nrDNA sequences of Suhomyces rilaensis DZ3T and related species using a neighbour-joining method (Kimura two-parameter model; MEGA v. 7; 100 bootstrap replicates). Dilnora E. Gouliamova & Roumen Dimitrov, The Stephan Angeloff Institute of Microbiology, Bulgarian Academy of Sciences, 26 Acad. Georgi Bonchev, Sofia 1113, Bulgaria; e-mail: dilnorag@gmail.com & roumen.dimitrov@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 392 Persoonia – Volume 45, 2020 Tolypocladium flavonigrum 393 Fungal Planet description sheets Fungal Planet 1175 – 19 December 2020 Tolypocladium flavonigrum Noisripoom, Tasanathai, Khonsanit & Luangsa-ard, sp. nov. Etymology. Named after the colour of fresh stromata, from the Latin ‘flavo’ meaning yellow, and ‘nigrum’ meaning black. The results of our phylogenetic study using LSU, tef1 and rpb1 sequences clearly separates T. flavonigrum from other species. Classification — Ophiocordycipitaceae, Hypocreales, Sordariomycetes. Based on a megablast search of NCBIs GenBank nucleotide database, the LSU sequence of Tolypocladium flavonigrum had the highest similarity to T. japonicum (strain OSC 110991, GenBank DQ51876.1; Identities = 850/908 (94 %), 33 gaps (3 %)), the closest hits using the tef1 sequence are T. japonicum (strain OSC 110991, GenBank DQ522330.1; Identities = 880 / 921 (96 %), no gaps), the closest hits using the rpb1 sequence are T. japonicum (strain OSC 110991, GenBank DQ522375.1; Identities = 534 /566 (94 %), 3 gaps (0 %)). Single or multiple stromata emerging directly from the ground growing on unidentified Elaphomyces sp., clavate, 15 – 30 mm long, 1– 2 mm wide, yellowish green when immature, black when mature. Terminal part of the stroma fertile, capitate, yellow black to black, 2 – 5 mm diam. Perithecia crowded, ordinal in arrangement, completely immersed, elongate-ovoid, (560 –) 567– 697(–750) × (200 –) 206 – 248 (– 250) µm. Asci cylindrical, 8-spored, (318 –) 330 – 416 (– 482) × 7– 8 µm with thickened ascus caps, 4.5 – 5 × 5 µm. Ascospores hyaline, filiform, (310 –) 330 – 375 (– 395) × 1.5 – 2 µm, breaking into 64 cylindrical part-spores, 2 – 5 × 1.5– 2 µm. Culture characteristics — (Colonies developed from germinating ascospores. Ascospores germinated within 24 h on potato dextrose agar (PDA). Colonies on PDA moderately growing, funiculose, c. 10 mm diam after 3 w at 25 °C. Colonies white to smoke grey with age. Colonies reverse cream. Conidiophores erect, arising from vegetative hyphae. Conidia single-celled, hyaline, smooth, globose, 3 – 5 µm diam, produced in slimy heads. Typus. thAilAND, Kalasin Province, Khok Pa Si Community Forest, on Elaphomyces sp., underground, 14 Aug. 2013, K. Tasanathai, W. Noisripoom & A. Khonsanit (holotype BBH37600, culture ex-type BCC66576 = MY08887, ITS, LSU and tef1 sequences GenBank MN338090, MN337287 and MN338495, MycoBank MB832658). Additional material examined. thAilAND, Kalasin Province, Khok Pa Si Community Forest, on Elaphomyces sp., underground, 14 Aug. 2013, K. Tasanathai, W. Noisripoom & A. Khonsanit, BBH37601, culture BCC66578, ITS, LSU and tef1 sequences GenBank MN338091, MN337288 and MN338496; ibid. BBH37602, culture BCC66580 LSU, tef1 and rpb1 sequences GenBank MN337289, MN338497 and MN338494. Notes — Tolypocladium flavonigrum is a rare species in Thailand found only in Khok Pa Si Community Forest, Kalasin province. Compared with other species occurring on Elaphomyces sp., T. flavonigrum shows similarity to T. fractum (Mains 1957) in the colour and shape of the fertile head as well as in the size and shape of the ascospores but differ in the size of perithecia and asci. Tolypocladium flavonigrum produces elongate, ovoid perithecia and asci, which are broader than T. fractum (500 – 600 × 220 – 260; 300 – 480 × 5 – 6 µm, respectively). Tolypocladium japonicum (Mains 1957) possesses a clavate fertile head and is distinct from all known species on truffles, while both T. capitatum (Mains 1957) and T. longisegmentum possess a yellowish stipe and reddish brown fertile part of the stromata, and differ mainly by the length of their part-spores: T. longisegmentum has the longest part-spores (40 – 65 × 4–5 µm) followed by T. capitatum, T. japonicum and T. flavonigrum (8–32 × 2.5–3; 10–18 × 2.5–4; 2–5 × 1.5–2 µm), respectively. Colour illustrations. Type locality – a small plot in Khok Pa Si Community Forest. Stroma on Elaphomyces sp.; immersed, elongate ovoid perithecia; part of asci showing asci tips; ascospore; part-spores; colonies on PDA; conidiophores with conidia; conidia. Scale bars =10 mm (stromata and plate culture), 100 µm (perithecia), 10 µm (asci and ascospore), 5 µm (part-spores, conidiophores with conidia and conidia). Tolypocladium flavonigrum BCC 66576 Tolypocladium flavonigrum BCC 66580 -/65/99 Tolypocladium flavonigrum BCC 66578 Tolypocladium japonicum OSC 110991 Tolypocladium fractum OSC 110990 Tolypocladium capitatum NBRC 100997 Tolypocladium longisegmentum OSC 110992 Tolypocladium capitatum OSC 71233 Tolypocladium fumosum WA 18945 Tolypocladium inegoense SU 15 Tolypocladium paradoxum Tolypocladium paradoxum NBRC 106958 90/94/100 Tolypocladium ophioglossoides OSC 106405 Tolypocladium ophioglossoides NBRC 8992 Tolypocladium cylindrosporum ARSEF 2920 Tolypocladium inflatum OSC 71235 59/81/100 Tolypocladium inflatum CBS 824.70 Tolypocladium cylindrosporum NRRL 28025 Tolypocladium album GB 5502 52/57/88 Tolypocladium album CBS 393.89 Tolypocladium endophyticum MX 575 100/99/100 Tolypocladium tropicale IQ 214 87/99/100 Tolypocladium tropicale MX 338 Tolypocladium amazonense MS 308 Tolypocladium geodes CBS 723.70 Tolypocladium pustulatum MF 5785 Tolypocladium jezoensis Purpureocillium lilacinum CBS 284.36 Purpureocillium lilacinum CBS 431.87 73/93/100 Drechmeria sinensis CBS 567.95 99/100/100 Drechmeria gunnii OSC 76404 96/100/95 Ophiocordyceps longissima EFCC 6814 62/74/- Ophiocordyceps longissima NBRC 108989 Ophiocordyceps sobolifera KEW 78842 Ophiocordyceps yakusimensis HMAS 199604 85/86/100 Ophiocordyceps sinensis EFCC 7287 Ophiocordyceps sinensis ARSEF 6282 97/93/100 Polycephalomyces formosus ARSEF 1424 Polycephalomyces ramosopulvinatus EFCC 5566 Polycephalomyces nipponicus BCC 2325 Cordyceps militaris OSC 93623 Cordyceps kyusyuensis EFCC 5886 65/100/100 Tolypocladium Purpureocillium Drechmeria Ophiocordyceps Polycephalomyces 50 changes Phylogenetic tree with T. flavonigrum constructed from a combined dataset comprising LSU, tef1 and rpb1. The phylogenetic tree was analysed using Maximum parsimony (MP), Maximum likelihood (ML) and Bayesian inference. The MP analysis was conducted on the combined data set using PAUP v. 4.0b10 (Swofford 2003), adopting random addition sequences (100 replications), with gaps being treated as missing data. A bootstrap (BP) analysis was performed using the maximum parsimony criterion in 1 000 replications. The ML analysis was run with RAxML-VI-HPC2 v. 8.2.12 (Stamatakis 2014) under a GTR model, with 1 000 bootstrap replicates. Bayesian phylogenetic inference was calculated with MrBayes v. 3.2.6 (Ronquist & Huelsenbeck 2003), with 5 M generations and under the same model. Numbers at the significant nodes represent MP bootstrap support values/RAxML bootstrap support values/Bayesian posterior probabilities (BPP) times 100. Thickened lines in the tree represent 99 –100 % bootstrap support values and 99 –100 BPP. Supplementary material FP1175 List of species and GenBank accessions numbers of sequences used in this study. Kanoksri Tasanathai, Wasana Noisripoom, Artit Khonsanit & Jennifer Luangsa-ard, Plant Microbe Interaction Research Team, Bioscience and Biotechnology for Agriculture, BIOTEC, 113 Thailand Science Park, Pathum Thani 12120, Thailand; e-mail: tasanatai@biotec.or.th, wasana.noi@biotec.or.th, Artit.kho@biotec.or.th & jajen@biotec.or.th © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 394 Persoonia – Volume 45, 2020 Tuber lusitanicum 395 Fungal Planet description sheets Fungal Planet 1176 – 19 December 2020 Tuber lusitanicum Ant. Rodr. & Muñoz-Mohedano, sp. nov. Etymology. Referring to Lusitania, the name given by the Romans to the western region of the Iberian Peninsula, which now covers the Portuguese area below Douro river and the neighbouring regions of Spanish Extremadura. Classification — Tuberaceae, Pezizales, Pezizomycetes. Ascomata hypogeous, 0.5 – 2 cm in size, subglobose, often lobed or irregular in form, solid, firm, white at first, becoming white-cream, pale yellowish, sometimes with a reddish tinge, darker at maturity, smooth. Peridium 300 – 500 μm thick, twolayered: the outermost pseudoparenchymatous, composed of subglobose or subangular cells, mostly 10 – 20 μm diam, yellowish, thick-walled, giving rise to hairs at the surface overlying; the inner layer composed of hyaline, thin-walled, interwoven, broad hyphae gradually intermixing into gleba. Hairs sparse, commonly 40–60 × 3–5 µm, hyaline, slender, tapered, setose, thick-walled, sometimes 1-septate near the base. Gleba whitish when immature, becoming olive brown, dark brown at maturity, marbled with numerous, thin, white veins, some veins ending in the peridium. Odour slight and not distinctive. Asci inamyloid, 50 – 80 × 50 – 60 μm, thin-walled, ellipsoid to subglobose, sessile or short-stalked, (1–)3 – 4(– 5)-spored. Ascospores 19 – 35 × 17– 28 μm, Q = 1.1–1.3, excluding ornamentation, the walls 2 µm thick, at first hyaline, becoming yellowish brown at maturity, subglobose to broadly ellipsoid, ornamented with a regular reticulum, alveoli 3 – 6 µm tall, 6 –10 µm long, 2 – 5 alveolar meshes along the spore length, polygonal (5–6 sides). Ecology & Distribution — Tuber lusitanicum grows in acidic soils of Extremadura dehesas associated to Quercus spp. in spring. Currently known only from Cáceres, Spain. Typus. spAiN, Cáceres, Rosalejo, in acidic soil, under Quercus suber (Fagaceae), 10 June 2012, A. Rodriguez (holotype MUB Fung-986, ITS and LSU sequences GenBank MT621651 and MT705332, MycoBank MB835881). Additional materials examined. spAiN, Cáceres, Rosalejo, under Quercus faginea, 10 June 2012, J. Mohedano, MUB Fung-987 and MUB Fung-988, ITS sequences GenBank MT621652 and MT621653; Belvís de Monroy under Quercus suber, 20 May 2012, J. Mohedano, MUB Fung-989 and MUB Fung990, ITS sequences GenBank MT621654 and MT621655; Millanes under Quercus suber, 5 June 2006, J. Mohedano, MUB Fung-991, ITS sequence GenBank MT621656. Notes — Tuber lusitanicum is a whitish truffle that clusters in the maculatum clade, and is characterised by its white-cream smooth peridium, brown gleba marbled with numerous, thin, white veins and reticulate-alveolate spores. Tuber lusitanicum is a sister species to T. rapaeodorum (88 % of similarity of ITS sequence), but T. rapaeodorum differs by having larger, narrower spores and thinner peridium (Ceruti et al. 2003). It also resembles T. maculatum (74 % of similarity of ITS sequence) but T. maculatum has a prosenchymatous peridium, lacking hairs and larger spores (Mello et al. 2000). DQ011845 100 JF261393 100 Tuber scruposum JF261405 100 JF261401 MT621657 100 91 MT621658 Tuber foetidum 100 AJ557544 KJ524540 100 Tuber maculatum 96 AJ557516 MT621651 holotype MT621652 100 MT621653 100 Tuber lusitanicum sp. nov. MT621654 MT621655 99 MT621656 AJ557521 EU784429 100 Tuber rapaeodorum DQ011849 100 DQ011850 KF744063 Tuber pseudosphaerosporum DQ898183 Tuber latisporum 0.10 Colour illustrations. Spain, Cáceres, Rosalejo, Quercus suber in Extremadura dehesa where the holotype was collected. Ascocarps; mature ascospores; peridium and hairs. Scale bars = 20 μm. Maximum likelihood (ML) phylogenetic tree inferred from ITS sequences using RAxML-HPC v. 8 (Stamatakis 2014) on XSEDE in the CIPRES science gateway (Miller et al. 2010). GTR + G was selected as model of evolution for the analysis. The sequences obtained in the present study are highlighted in bold. Bootstrap support values (≥ 70 %) are indicated at the nodes. Tuber latisporum and Tuber pseudosphaerosporum were used as outgroup. The scale bar indicates the expected changes per site. Antonio Rodríguez, Justo Muñoz-Mohedano, Alfonso Navarro-Ródenas, Francisco Arenas & Asunción Morte, Departamento de Biología Vegetal (Botánica), Facultad de Biología, Universidad de Murcia, 30100 Murcia, Spain; e-mail: antonio@trufamania.com, kaerques@gmail.com, anr@um.es, f.arenasjimenez@um.es & amorte@um.es © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 396 Persoonia – Volume 45, 2020 Tylopilus subotsuensis 397 Fungal Planet description sheets Fungal Planet 1177 – 19 December 2020 Tylopilus subotsuensis T.H.G. Pham, A.V. Alexandrova & O.V. Morozova, sp. nov. Etymology. The epithet refers to macromorphological similarity of the new species to Tylopilus otsuensis. Classification — Boletaceae, Boletales, Agaricomycetes. Basidiomata medium to large sized, boletoid. Pileus 30‒90 mm diam, firstly hemispherical, then convex and pulvinate-flattened; fleshy; margin initially involute, then curved downwards, finally plane, not or only slightly extending beyond the tubes, surface matt, dry, slightly slimy in moist weather, firstly finely pruinose or felted, then smooth and glabrous; the colour varies from beige and pale ochraceous brown with minute olive tinge to light brown and brown (4A3‒4, 4B3‒4, 5C3‒5, 5D4‒8, 6E4‒6; Kornerup & Wanscher 1978). Hymenophore depressed around the apex of stipe, up to 10 mm thick, thinner than the context, whitish, becoming pinkish; pores round or slightly angular, up to 1 mm diam. Stipe 80 –120 × 10 – 25 mm, almost cylindrical or broadened towards the base, solid; minutely tomentose, without distinct reticulum; dry, slightly slimy in moist weather, concolorous with the pileus or slightly lighter. Context firm, white, unchanging or slowly yellowing in the stem base and in the areas damaged by insects. Smell weak, taste bitter. Spores (7.5 –) 9 –10(–11.5) × (3 –)3.5(– 4.5) μm, Q = (2.2 –)2.6(– 3), fusoid, ellipsoid-fusoid, tapering towards the apex, inequilateral in side view, without or with weak suprahilar depression, sometimes substrangulate, hyaline in KOH, smooth. Basidia 23–28 × 6–9 μm, 4-spored, narrowly clavate to clavate, clampless. Cheilocystidia lageniform or fusoid, 29 – 66 × 9 –15 μm, often thin-walled, with granulose content, forming sterile or heterogeneous tube edge. Pleurocystidia 45 –72 × 10 –15 μm, same as cheilocystidia. Hymenophoral trama divergent, boletoid. Pileipellis a trichoderm, made up of strongly interwoven filamentous, frequently branched yellowish hyphae 5 – 8 μm wide. Stipitipellis a caulohymenium of basidiolae-like narrowly clavate cells, 25 – 35 × 7–10 μm, with scattered caulobasidia. Caulocystidia 25 – 60 × 6 –10 μm, lageniform, fusoid or subcylindrical. Clamp connections absent. Habit, Habitat & Known distribution — Solitary, in groups or caespitose on soil in montane evergreen tropical forests. Known from Vietnam. Additional materials examined. VietNAm, Dak Lak Province, Krong Bong District, Chu Yang Sin National Park, Krong Kmar, 7 km northwest of Chu Yang Sin mountain, N12.39497° E108.34823°, 1 000 m alt., middle montane evergreen mixed riparian forest, 21 Mar. 2013, A.V. Alexandrova & T.H.G. Pham (LE312526; tef1α sequence GenBank MW014271); ibid., 22 Mar. 2013, A.V. Alexandrova & T.H.G. Pham (LE312525; ITS and tef1α sequences GenBank MW009075 and MW014269); Lam Dong Province, Bao Lam District, 21 km NW of the town of Bao Loc, Loc Bac Forestry, N11.74449° E107.70647°, 1 006 m alt., 6 Apr. 2013, lower montane evergreen broadleaf forest (Magnoliaceae, Myrtaceae, Theaceae, Lauraceae, Fagaceae, Annonaceae), A.V. Alexandrova & T.H.G. Pham (LE312528; tef1α sequence GenBank MW014270); Gia Lai Province, K’Bang District, Son Lang Commune, Kon Chu Rang Nature Reserve, N14.50042° E108.56338°, 1 000 m alt., on soil in middle montane evergreen mixed forest, 27 May 2016, A.V. Alexandrova (LE312527; tef1α sequence GenBank MW014272). Notes — Tylopilus subotsuensis is characterised by the brownish basidiomata, usually lacking distinct olivaceous or purplish tinges. Tylopilus otsuensis, described from Japan (Hongo 1966), is superficially similar. It is distinguished by the oblong spores, presence of distinct olivaceous tinge and reddish-brown discolouration when bruised. Long, fusoid, tapering towards the apex spores, unchanging or slightly yellowish context and lack or almost lack of olivaceous tinge in the colour of basidiomata are characteristic for the new species. Micromorphologically, due to spores and cystidia, the new species resembles T. neofelleus (Hongo 1973). But for the latter species the presence of a more or less pronounced purplish tinge and thin reticulum on the upper part of stipe surface is characteristic. In fact, colour variations are not a very reliable way to distinguish species in the genus Tylopilus. Gelardi et al. (2014b) and Wu et al. (2016) have shown with molecular evidence that the presence of a purplish tinge in basidiomata of T. neofelleus (including T. microsporus) can vary greatly from a pronounced colour to its complete absence. In the last case T. subotsuensis and T. neofelleus are almost inseparable, distinguished only by the absence of a reticulum in the apex of the stipe, slightly longer spores and wider cystidia in the new species. However, molecular data support the T. subotsuensis as distinct. Typus.VietNAm, Dak Lak Province, Krong Bong District, Chu Yang Sin National Park, Krong Kmar, 7 km northwest of Chu Yang Sin Mt., N12.42656° E108.36633°, 985 m alt., middle montane evergreen broadleaf forest, 18 May 2014, A.V. Alexandrova & T.H.G. Pham (holotype LE312534; ITS, tef1α and LSU sequences GenBank MW009074, MW014268 and MW009073, MycoBank MB837493). Colour illustrations. Vietnam, Dak Lak Province, Krong Bong District, Chu Yang Sin National Park, type locality. Spores, cheilocystidium; pleurocystidium; pileipellis; stipitipellis with caulocystidia (all from holotype). Pileus, basidioma in situ (from holotype); group of fasciculate basidiomata with a longitudinal section through one of them. Scale bars = 10 µm (spores and microstructures), 1 cm (basidiomata). Supplementary material FP1177 Phylogenetic tree derived from Bayesian analysis based on tef1α data. The analysis was performed under a GTR model of evolution for 3 M generations using MrBayes v. 3.2.1 (Ronquist et al. 2012). Posterior probability (PP > 0.95) values from the Bayesian analysis are shown at the nodes. The scale bar represents the expected number of nucleotide changes per site. Thi Ha Giang Pham, Joint Russian-Vietnamese Tropical Research and Technological Center, Hanoi, Vietnam; e-mail: giangvietnga@gmail.com Alina V. Alexandrova, Lomonosov Moscow State University (MSU), Faculty of Biology, 119234, 1, 12 Leninskie Gory Str., Moscow, Russia / Joint Russian-Vietnamese Tropical Research and Technological Center, Hanoi, Vietnam; e-mail: alexandrova@mail.bio.msu.ru Olga V. Morozova, Komarov Botanical Institute of the Russian Academy of Sciences, 197376, 2 Prof. Popov Str., Saint Petersburg, Russia; e-mail: OMorozova@binran.ru © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 398 Persoonia – Volume 45, 2020 Veloboletus limbatus 399 Fungal Planet description sheets Fungal Planet 1178 – 19 December 2020 Veloboletus limbatus Fechner & Halling, gen. & sp. nov. Etymology. Velo- (veil) + boletus (genus of Boletaceae); limbus- (edge, rim, margin referring to the obvious veil remnant edge at the base of stipe). Classification — Boletaceae, Boletales, Agaricomycetes. Pileus (3–)6.5–9 cm broad, convex to plano-convex, dry, finely appressed squamulose, with squamules brown to deep reddish brown or reddish brown, overlying a dull red disc, dull yellow to bright yellow or dirty yellowish at margin, staining blue, with even or rarely a sterile projecting margin attached at stipe when young to form limbate rim on stipe base. Flesh pale yellow to pale lemon yellow, 1– 2 cm thick, staining blue when exposed, with mild odour and slowly unpleasant to nearly bitter taste. Hymenophore adnexed to depressed around stipe, with tubes bright yellow to bright greenish yellow (2A– B7,6; Kornerup & Wanscher 1983), sometimes hardly bluing when young or staining blue-green when bruised at first, with pores olive yellow (3C8) to olive brown, then pores becoming brown. Stipe (4.5 –) 6.5 – 9.5 cm long, (1.5 –)2 – 2.7(– 5) cm broad, dry, terete or slightly flattened, equal to subclavate to clavate, sometimes with a pinched base, with a clearly defined limbate rim and tapering to base below that; surface bright lemon yellow above and heavily pruinose to subfloccose or fibrillose streaked, matted toward base, fading to whitish with age, pale pinkish red with fine appressed to suberect deep red to brown squamules below limbate rim (as in pileus), staining blue; interior solid, yellow, staining blue, with yellow (4B8) basal mycelium. Basidiospores 9.6 –15.2 × 3.5 – 4.9 μm, x = 12.43 × 4.30 μm, Q = 2.89, n = 100, p = 5, smooth, subfusoid to ellipsoid, hyaline to pale yellow in KOH, hyaline to rarely weakly dextrinoid in Melzer’s. Basidia 18 – 40 × 8 –12 μm, 4-sterigmate, clavate, hyaline, inamyloid. Pileus trama interwoven with hyaline, thinwalled hyphae, 4 –15.6(– 20) µm broad. Tube trama boletoid and divergent, becoming gelatinised with age, with hyphae 4 –15.6(– 20) µm broad, hyaline in KOH and Melzer’s. Pleurocystidia 32 – 41.6 × 8.8 –12 µm clavate, thin walled, inamyloid. Cheilocystidia 8 – 34.4 × 6.4 –10.4 μm obclavate to clavate, inamyloid, thin-walled. Pileipellis a trichodermium, composed of erect to suberect cylindrical elements, 3.2 – 9.6 μm broad, smooth, thin-walled, hyaline to occasionally very slightly dextrinoid. Stipitipellis a fragile and indistinct layer of cylindric to clavate elements, 3.2 –12 μm long, smooth, thin-walled, inamyloid, hyaline. Clamp connections absent. Habitat & Distribution — Solitary to gregarious on soil or sand under Allocasuarina sp., Eucalyptus sp., and Eucalyptus grandis. At present, known in Queensland from the Tablelands west of Cairns southward to Fraser Island and the southern border of the state in the mountains west of the Gold Coast. In the months February to March, June. Colour illustrations. Sclerophyll vegetation with Eucalyptus and Allocasuarina at Camp Milo of the Cooloola Sandmass near Fraser Island. Stipitipellis; basidiospores; pileipellis; holotype (REH9228); solitary basidiome (REH8746); sectioned basidiome (REH8917) showing universal veil attachment (arrows). Scale bars = 1 cm (entire basidiomes), 0.5 cm (sectioned basidiome); 10 µm (spores and stipitipellis), 40 µm (pileipellis). Typus. AustrAliA, Queensland, Wide Bay District, Great Sandy National Park, Fraser Island, Kingfisher Bay, S25°23'35.7" E153°1'50.7", 8 m, 10 June 2009, R.E. Halling 9228 (holotype BRI AQ0794331, isotype NY 1393645; rpb2, atp6, tef1 and LSU sequences GenBank MT747397, MT747398, MN413636 and MN393700, MycoBank MB832369 (genus), MB832370 (species)). Notes — BLAST searches were conducted against the NCBIs GenBank nucleotide database for each of the six novel sequences using megablast in the blastn suite (Johnson et al. 2008). The results based on percent identity indicated consistent placement of Veloboletus limbatus in the subfamily Xerocomoideae (family Boletaceae). This was corroborated by a series of phylogenetic analyses of individual genes with selections of exemplars from across the Boletaceae (especially Xerocomoideae), Paxillaceae, and Suillaceae. A concatenated analysis of tef1 and LSU was also done in this manner (see Supplement material FP1178). These analyses were conducted with MrBayes v. 3.2.7A (Ronquist et al. 2012) on the CIPRES REST API (Miller et al. 2015). In all cases, Veloboletus limbatus was consistently placed within a highly-supported Xerocomoideae (Bayesian posterior probability (bpp) = 1). With tef1 and LSU, where more than one specimen of V. limbatus was available, the genus was fully supported (bpp = 1) in the individual and concatenated analyses. Though we were able to infer that Veloboletus belongs within subfamily Xerocomoideae, no clear sister group to Veloboletus was apparent. As far as we know, there are no other members of the Boletaceae with a distinctive and conspicuous squamulose, universal veil rupturing to form an obvious limbate rim. That and the conspicuous cyanescence are diagnostic. Xerocomoideae is globally diverse and contains a number of iconic mushroom groups, including for example, Boletellus, Aureoboletus, Phylloporus, Pulchroboletus, Heimioporus, and Xerocomus s.str. It is notable that Veloboletus limbatus has a universal veil. According to the terminology of Clémençon (2012), V. limbatus exhibits a cleistometablema. Several other epigeous stipitatepileate Xerocomoideae exhibit veils that could be interpreted as universal (Boletellus ananas, B. ananiceps, B. emodensis, B. deceptivus, B. singeri, Aureoboletus longicollis), but in those species, the portion of the veil nearest the stipe is not physically connected to the stipe tissue. In B. singeri and A. longicollis, the veil can separate from the pileus margin and form an annular appendage. Alessioporus ichnusanus, a Xerocomoideae from southern Europe, is described as leaving a velar remnant on the stipe of mature fruiting bodies due to mixangiocarpic development (Gelardi et al. 2014a). Based on the combination of morphological features alone, we hypothesize the uniqueness of the taxon merits generic recognition. Clearly, the need for further exploration and collection of Boletales in Australasia, which harbours a diverse and unique mycota, is required. Future fieldwork or herbarium-based studies may uncover a sister group to Veloboletus or reveal additional species in the genus. Supplementary material FP1178-1 Additional materials examined. FP1178-2 Bayesian phylogram of selected Boletales, especially subfam. Xerocomoideae. Roy E. Halling, Institute of Systematic Botany, New York Botanical Garden, 2900 Southern Blvd, Bronx, NY 10458-5126, USA; e-mail: rhalling@nybg.org Nigel Fechner, Queensland Herbarium, Mt Coot-tha Road, Toowong, Brisbane, Queensland 4066, Australia; e-mail: nigel.fechner@des.qld.gov.au Naveed Davoodian, National Herbarium, Royal Botanic Gardens Victoria, South Yarra, Victoria 3141, Australia; e-mail: naveed.davoodian@rbg.vic.gov.au © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 400 Persoonia – Volume 45, 2020 Xenomonodictys iranica Fungal Planet description sheets 401 Fungal Planet 1179 – 19 December 2020 Xenomonodictys Hern.-Restr., Karimi, Alizadeh & Tajick Ghanbary, gen. nov. Etymology. From the Greek ‘Xenos’ indicating strangeness and the related genus Monodictys, referring to a variant of the genus Monodictys. Classification — Sporormiaceae, Pleosporales, Dothideomycetes. Conidiophores micronematous, hyaline to subhyaline, mostly reduced to conidiogenous cells arising from the mycelium. Conidiogenous cells hyaline to subhyaline, cylindrical. Conidia multicellular, composed by brown cells of different tones, usually with basal cell paler than the rest. Type species. Xenomonodictys iranica Hern.-Restr., Karimi, Alizadeh & Tajick Ghanbary. MycoBank MB837750. Xenomonodictys iranica Hern.-Restr., Karimi, Alizadeh & Tajick Ghanbary, sp. nov. Etymology. Name refers to the country where this fungus was collected, Iran. Mycelium composed of hyaline to brown, smooth, septate, 1– 2 µm wide hyphae. Conidiophores micronematous, hyaline to subhyaline, mostly reduced to conidiogenous cells arising from the mycelium. Conidiogenous cells hyaline to subhyaline, cylindrical, 3 – 9 × 1– 2 µm. Conidia 10 –15 × 7– 9 µm, base 1– 2 µm, subglobose to ellipsoidal, multicellular, composed of up to 10, smooth, brown cells, each cell 3 – 5 µm diam, usually in two rows, with basal cell paler than the rest. Conidial secession rhexolythic. Culture characteristics — Colonies in oatmeal agar (OA) at 25 °C reaching 40 mm after 3 wk, cottony to velvety, with moderate aerial mycelium, olivaceous grey to mouse grey, margin entire to fimbriate; reverse olivaceous grey. On potato dextrose agar (PDA) after 2 wk reaching 45 mm, greyish to black. Typus. irAN, Mazandaran, Pol sefid, (N36°3'27.99" E53°5'57.84"), on wood of Fagus orientalis (Fagaceae), 11 May 2015, O. Karimi A2FC200 (holotype CBS H-24521, culture ex-type CBS 147181, ITS and LSU sequences GenBank MW175368.1 and MW175406.1, MycoBank MB837751). Notes — Monodictys is a large genus with 69 names presently registered in Index Fungorum. Morphologically it is characterised by multicellular, brown conidia borne on micronematous conidiophores. Based on DNA sequence data, species of Monodictys have in the past been allocated to several genera in Dothideomycetes and Sordariomycetes. Monodictys putredinis, the type species, is the asexual morph of Ohleria brasilensis (Melanommataceae) which resides in Pleosporales together with Paramonodictys (Parabambusicolaceae), Pleomonodictys (Pleomonodictydaceae) and Xenomonodictys (Sporormiaceae). Other species (as asexual morphs) have been connected with Tubeufia (Tubeufiaceae) and Aquastroma (Parabambusicolaceae) in Dothideomycetes (Day et al. 2006, Velmurugan et al. 2013, Tanaka et al. 2015, Hernández-Restrepo et al. 2017, Vu et al. 2019). In Sordariomycetes, however, monodictys-like species are placed in the genera Dematiosporium, Ascotaiwania (Savoryellalceae), Neomonodictys (Pleurotheciaceae), Trichocladium (Chaetomiaceae), and Nereiospora (Microascales) (Mouzouras & Jones 1985, Hernández-Restrepo et al. 2017, Réblová et al. 2020). Furthermore, in Helotiales, a monodictys-like species has been accommodated as the asexual morph of Hyaloshypha monodictys (Hosoya & Huhtinen 2002). Xenomonodictys is therefore introduced as a new genus for a monodictys-like taxon phylogenetically related to Preussia terricola. Based on a megablast search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to Pleospora iqbalii (GenBank NR_160118.1; Identities = 486/546 (89 %), 28 gaps (5 %)), and Preussia fleischhakii (GenBank MH474379.1; Identities = 484/549 (88 %), 18 gaps (3 %)). Closest hits using the LSU sequence are Preussia terricola (GenBank GQ203725.1; Identities = 820/845 (97 %), one gap (0 %)), Pleospora iqbalii (GenBank MH871062.1; Identities = 819/847 (97 %), five gaps (0 %)), and Neomassarina chromolaenae (GenBank NG_068715.1; Identities = 817/845 (97 %), two gaps (0 %)). Colour illustrations. Farim Forest near the city of Pol Sefid, Mazandaran Province, Iran. Conidiophores and conidia; conidia. Scale bars = 20 μm (conidiophores), 10 μm (all others). Margarita Hernández-Restrepo, Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508 AD Utrecht, The Netherlands; e-mail: m.hernandez@wi.knaw.nl Omid Karimi & Mohammad Ali Tajick Ghanbary, Department of Plant Protection, Faculty of Agronomy, Sari Agricultural Sciences and Natural Resources University, Sari, Iran; e-mail: karimiomid18@gmail.com & mycology2@gmail.com Alireza Alizadeh, Department of Plant Protection, Faculty of Agriculture, Azarbaijan Shahid madani University, Tabriz, Iran; e-mail: alizadeh.al2008@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 402 Persoonia – Volume 45, 2020 Cortinarius indopurpurascens 403 Fungal Planet description sheets Fungal Planet 1180 – 19 December 2020 Cortinarius indopurpurascens Dima, Semwal, Brandrud, V. Papp, & V.K. Bhatt, sp. nov. Etymology. The epithet refers to the occurrence in India and the close relationship with Cortinarius purpurascens. Classification — Cortinariaceae, Agaricales, Agaricomycetes. Pileus 45 – 65 mm diam, convex to plano-convex, then applanate, margin uplifted with age, surface sticky to glutinous, glabrous, with a few darker, hygrophanous spots or radial streaks, initially pale bluish grey (Methuen 12B3–12C3) with an ochraceous tinge at disc, then becoming somewhat ochraceous brown from centre (6C6, 6D8 – 6C5). Lamellae emarginate, crowded, bifurcate towards margin, up to 7 mm broad, lamellulae of various lengths, bright purple, amethyst to reddish lilac tinge (15A6, 15C5 –14B5), turning darker purplish when bruised. Stipe 50 –70 × 10 –17 mm, cylindrical with a 17– 24 mm wide roundish marginated bulb at the base, concolorous with lamellae, purple to amethyst (15A6 –15C5), reddish lilac tinged (14B5) when mature or bruised. Few remnants of cortina present at the upper half. Universal veil at bulb margin thin and indistinct. Context purplish. Odour honey-like, especially when bruised. Taste not recorded. Spore print cocoa brown (6E7). Basidiospores (9.3 –)9.7–10.3(–10.9) × (5.2 –) 5.4 – 5.7(– 5.9) μm, av. = 9.9 × 5.5 μm, Q = (1.6 –)1.7–1.8(–1.9), Qav = 1.75, n = 60, (ellipsoid to) subamygdaloid, strongly verrucose, with discrete, hardly interconnected warts. Basidia 4-spored, 26–33 × 6– 8 μm, clavate. Habitat & Distribution — Solitary to caespitose, occurring among leaf litter of the evergreen banj oak Quercus leucotrichophora, on humicolous soil, in temperate broadleaved, Himalayan mid-elevation forests, dominated by mainly Q. leucotrichophora, Myrica esculenta with scattered Rhododendron arboreum trees. Typus. iNDiA, Uttarakhand, Pauri Garhwal, Mundneshwar, 1820 m asl, N29°01'5" E78°44'32", 12 Aug. 2015, K.C. Semwal (holotype KCS 2442, ITS sequence GenBank MW135432, MycoBank MB837766). Additional materials examined. iNDiA, Uttarakhand, Pauri Garhwal, Phedhkal, 1880 m asl, N30°16'36" E78°85'42", 17 Aug. 2015, K.C. Semwal, KCS 2467, ITS sequence GenBank MW135431; Dandapani, 1900 m asl, 28 July 2015, K.C. Semwal, KCS 2529, ITS sequence GenBank MW135430. Notes — Cortinarius indopurpurascens belongs to the sect. Purpurascentes based on morphological and molecular (nrDNA ITS and LSU regions) data. The species in this section are characterised by basidiomata with purplish lilac tinges mainly in young stages of development, surfaces and context becoming purplish-lilac on bruising, especially on the lamellae, a positive Lugol reaction in the context, moderate to strong honey-like smell, and ellipsoid to subamygdaloid, distinctly-strongly verrucose spores (Saar et al. 2014, Soop et al. 2019). The nrDNA ITS sequences of the three studied C. indopurpurascens specimens are identical and form a well-supported monophyletic group within sect. Purpurascentes closely related to the European C. purpurascens, and to an undescribed Cortinarius species from North America (see Supplementary Material FP1180). It differs by 9 –10 nucleotide and indel positions (98.5 – 98.3 % similarity) from C. purpurascens and 6 – 8 nucleotide and indel position (99 – 98.7 % similarity) from Cortinarius sp. In morphology C. indopurpurascens is most similar to the mainly European C. purpurascens and C. collocandoides; both species may possess strong lilac-purplish tinges on the basidiomata, and show a distinctly marginate bulb. According to material seen, C. indopurpurascens seems to be a paler species, being pale bluish grey when young, a colour reminding more of C. porphyropus (a more distant relative with non-marginated bulb), than of C. purpurascens. With regard to microcharacters, the European species have significantly smaller spores (C. purpurascens: av. = 8 × 4.9 μm and C. collocandoides: av. = 9.2 × 5.4 μm vs C. indopurpurascens: 9.9 × 5.5 μm). Furthermore, among the five European species in this section (Saar et al. 2014), all taxa have smaller and broader spores than those of C. indopurpurascens. Ecologically, C. indopurpurascens seems to be associated with the Himalayan, evergreen Quercus leucotrichophora, whereas, the closely related, mainly European C. purpurascens is associated with a wide range of trees, including oaks, but normally do not occur under (Mediterranean) evergreen oaks. It should be noted that C. purpurascens also follows the coniferous boreal-taiga belt into Asian Siberia, but here it is known only from Pinus sylvestris forests (pers. obs.), and it is highly unlikely that C. indopurpurascens and C. purpurascens have an overlapping distribution. Supplementary material Colour illustrations. India, Uttarakhand, Pauri Garhwal, Mundneshwar, type locality. Spores and basidiomata (from KCS 2442, holotype). Scale bar = 10 µm (spores). FP1180 Phylogenetic tree of Cortinarius sect. Purpurascentes derived from Maximum Likelihood analysis based on nrITS1-5.8S-ITS2 and binary data from indel coding with FastGap v. 1.2 (Borchsenius 2009). Analysis was performed in raxmlGUI v. 1.5.2 (Silvestro & Michalak 2012) using the GTRGAMMA substitution model for the partitioned (ITS1-5.8S-ITS2) nucleotide data and the default setting for binary (indel) data. ML bootstrap support (BS) values are shown at the nodes (BS > 70 %). Sequences of the new species are highlighted in blue. Bálint Dima, Department of Plant Anatomy, Institute of Biology, Eötvös Loránd University, Pázmány Péter sétány 1/C, H-1117, Budapest, Hungary; e-mail: cortinarius1@gmail.com Kamal C. Semwal, Department of Biology, College of Sciences, Eritrea Institute of Technology, Mai Nafhi, Asmara, Eritrea; e-mail: kamalsemwal@gmail.com Tor Erik Brandrud, Norwegian Institute for Nature Research, Gaustadalléen 21, NO-0349 Oslo, Norway; e-mail: tor.brandrud@nina.no Viktor Papp, Institute of Horticultural Plant Biology, Szent István University, H-1518, Budapest, Hungary; e-mail: agaricum@gmail.com Vinod K. Bhatt, Navdanya, 105, Rajpur Road, Dehradun, Uttarakhand, India; e-mail: vinodkbhatt@gmail.com © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 404 Persoonia – Volume 45, 2020 Cortinarius glaucoelotus Fungal Planet description sheets 405 Fungal Planet 1181 – 19 December 2020 Cortinarius glaucoelotus Brandrud, Dima, Krisai, Ballarà & Peintner, sp. nov. Etymology. Name refers to bluish (glaucous) tinges on stipe and resemblance to C. elotus sensu Moser. some Picea orientalis) and Altai (Russian Siberia; with Picea obovata). Classification — Cortinariaceae, Agaricales, Agaricomycetes. Typus. AustriA, Lower Austria, Schneebergdörfl , 780 m asl, N47°46'45'' E15°52'13'', 9 Oct. 2017, T.E. Brandrud, I. Krisai-Greilhuber & H. Voglmayr (holotype TEB898-17 (O), isotypus WU 42513, ITS sequence GenBank MW135358, MycoBank MB837764). Pileus 35 – 60 (– 80) mm diam, hemispherical, then planoconvex, glutinous, glabrous, often with coarse, radial innately fibrillose structure near margin, centre sometimes with whitish patches of universal veil remnants; initially olivaceous brown, often with greenish grey tinge at margin, more ochraceous brown-buff to brown at centre, exposed parts often becoming oxidised to warmer red-brown, even chestnut brown with age. Lamellae emarginate, 4 – 8 mm broad, greyish to faintly wax yellow-ochre tinged, paler towards margin, later greyish brown, edge even to slightly serrulate. Stipe 25 – 60 × 10 –17 mm, with a distinctly (but not very broad) marginate bulb (bulb up to 27 mm wide), whitish, with a distinct and sometimes persistent lilac-amethyst zone at apex, with age brownish. Universal veil on bulb margin sparse, viscid, often difficult to distinguish from stipe surface, whitish (bluish tinges not seen), cortina abundant, whitish, soon brown from spores. Context whitish to cream, some with pale ochre grey hygrophanous streaks at stipe apex, cortex lilac at stipe apex, a few with vivid saffron yellow spots in base of bulb. Odour distinctly raphanoid (to earthy). Taste mild. Spore print dark (rusty) brown. Basidiospores (9.8 –)10.9 –11.9(–12.7) × (6.4 –)6.7–7.5(–7.8) μm, av. = 11.40 × 7.07 μm, Q = (1.4–)1.5 –1.7(–1.8), Qav = 1.61 (type collection; n = 67); range of MVs from all collections 11.4 –12.1 × 7.0–7.7 μm, av. = 11.81 × 7.37 μm, Qav = 1.60; distinctly citriform to amygdaloid, strongly and coarsely, net-like verrucose, suprahilar plage indistinct, apiculus smooth. Basidia 4-spored, 9 –11 μm wide. Pileipellis simplex, with gradual transition from erect-entangled-sinuous, gelatinous, very narrow, 2–3 μm wide, pale yellow hyphae at surface, to more repent-parallel, slightly wider (3 – 4(– 5)) μm hyphae basally. The basal epicutis with pale yellow brown hyphae, sometimes more strongly yellow to yellow brown, with some hyphae filled with amorphousoleiferous golden brown pigment, pigment sometimes in lumps like staples of coins, hyphae sometimes forming subparallel, interconnected bundles. A few thicker (6 –7 μm wide) hyphae with faintly thicker walls are sometimes seen basally, some of these might be distinctly zebra-striped encrusted. Chemical reactions — KOH 20 – 30 % negative (slightly brownish) on pileipellis and bulb margin. Habitat & Distribution — In calcareous Abies dominated forests, as well as calcareous Picea and Pinus forests. Very rare, but widely distributed in montane Europe (Pyrenees - The Alps - Caucasus), and into Asian Siberia. Result of nrDNA ITS sequencing verified the species from NE Spain (with Pinus nigra and P. sylvestris), E Austria (with Abies alba and some Picea abies), Russian W Caucasus (with Abies nordmanniana and Colour illustrations. Austria, Schneebergdörfl, WU 42513, type locality. Spores and basidiomata (from WU 42513, holotype). Scale bar = 10 µm (spores). Additional materials examined. AustriA, Lower Austria, Schneebergdörfl NW, 9 Oct. 2017, T.E. Brandrud, I. Krisai-Greilhuber & H. Voglmayr, WU 42455 / TEB898b-17 (O), ITS sequence GenBank MW135357. – RussiA, Altai republic, Chuya river (Katun), NW of Uagan Unus, 22 Aug. 2001 (as ‘C. elotus’), M.M. Moser (IB 2001/0090), ITS-LSU sequence GenBank EU056953; Karachay-Cherkessia Republic (NW Caucasus), Kuzgych river, Arkhyz W, 10 Oct. 2016, T.E. Brandrud & T. Svetasheva, TEB 635-16 (LE), ITS sequence GenBank MW135356. – spAiN, Berguedà, Espunyola Can Gomira, 700–900 m asl, 8 Nov. 2014, J. Ballarà, JB-8525-14, ITS sequence GenBank MW135354. Notes — Cortinarius glaucoelotus belongs to the Humolentes clade (within the Calochroi lineage), where it has a sister position to C. pseudoglaucopus and C. praetermissus. Cortinarius pseudoglaucopus is distributed both in Europe and North America, and shows a slight phylogeographical differentiation between these regions. The nrDNA ITS sequences generated from C. glaucoelotus differ from sequences of European C. pseudoglaucopus by 9 –12 nucleotide and indel positions (98.35 – 98.02 % similarity). The two species have overlap in their distribution and similar habitat requirements: both are occurring in calcareous coniferous forests, but C. glaucoelotus is possibly more associated with Abies spp. They are co-occurring in W Caucasus and found in the same area and same kind of forests in E Austria and NE Spain. Morphologically, C. glaucoelotus and C. pseudoglaucopus are very similar, characterised e.g. by their initially olive brown pilei. Based on the material seen so far, it seems that C. glaucoelotus can be distinguished from C. pseudoglaucopus by the beautiful lilac-amethyst narrow zone at the stipe apex when young. However, we do not know if this lilac apex colour is constant. The bluish pigments in C. pseudoglaucopus are very variable, sometimes the stipe and context can be bluish-violet tinged when young, and the veil and bulb margin are often violaceous spotted when young. These bluish variants have usually rather dark pileus colours, as already indicated by Moser (1961). The bluish tinges on the bulb margin including the veil have not thus far been seen in C. glaucoelotus. It is also possible that C. glaucoelotus on average has more olive greenish tinges on the pileus margin when (very) young, but this needs further confirmation. (text continues on Supplementary material page FP1181) Supplementary material FP1181 Phylogenetic tree of the Humolentes clade within sect. Calochroi derived from a Maximum Likelihood analysis based on nrITS1-5.8S-ITS2, partial nrLSU and binary data from indel coding with FastGap v. 1.2 (Borchsenius 2009). Analysis was performed in raxmlGUI v. 1.5.2 (Silvestro & Michalak 2012) using the GTRGAMMA substitution model for the partitioned (ITS1-5.8S-ITS2 and LSU) nucleotide data and the default setting for binary (indel) data. ML bootstrap support (BS) values are shown at the nodes (BS > 70 %). Sequences generated for this study are highlighted in bold face. Bálint Dima, Department of Plant Anatomy, Institute of Biology, Eötvös Loránd University, Pázmány Péter sétány 1/C, H-1117, Budapest, Hungary; e-mail: cortinarius1@gmail.com Tor Erik Brandrud, Norwegian Institute for Nature Research, Gaustadalléen 21, NO-0349 Oslo, Norway; e-mail: tor.brandrud@nina.no Irmgard Krisai-Greilhuber, Department of Botany and Biodiversity Research, University of Vienna, Rennweg 14, 1030 Wien, Austria; e-mail: irmgard.greilhuber@univie.ac.at Josep Ballarà, C/ Tossalet de les Forques, 44, E-08600, Berga, Catalonia, Spain; e-mail: josep.cortinarius@gmail.com Ursula Peintner, Institute of Microbiology, University of Innsbruck, Technikerstrasse 25, 6020 Innsbruck, Austria; e-mail: Ursula.Peintner@uibk.ac.at © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 406 Persoonia – Volume 45, 2020 REFERENCES Anisimova M, Manuel G., Dufayard J-F, et al. 2011. Survey of branch support methods demonstrates accuracy, power, and robustness of Fast Likelihoodbased approximation schemes. Systematic Biology 60: 685 – 699. Antonín V, Ryoo R, Ka K-H, et al. 2014. Three new species of Crinipellis and one new variety of Moniliophthora (Basidiomycota, Marasmiaceae) described from the Republic of Korea. Phytotaxa 170: 86 –102. Arauzo S, Iglesias P. 2014. La familia Geoglossaceae ss. str. en la península Ibérica y la Macaronesia. Errotari 11: 166 – 259. Attili-Angelis D, Duarte APM, Pagnocca FC, et al. 2014. Novel Phialophora species from leaf-cutting ants (tribe Attini). Fungal Diversity 65: 65 –75. Australian Plant Pest Database, online database, (accessed 06 August 2020). https://appd.ala.org.au/appd-hub/search#tab_simpleSearch. Baral H-O. 1987. Lugol’s solution / IKI versus Melzer’s reagent: hemiamyloidity, a universal feature of the ascus wall. Mycotaxon 29: 399 – 450. Basiewicz M, Weiss M, Kogel K-H, et al. 2012. Molecular and phenotypic characterization of Sebacina vermifera strains associated with orchids, and the description of Pirformospora williamsii sp. nov. Fungal Biology 116: 204 – 213. Baten MA, Mingzhu L, Motohashi K, et al. 2015. Two new species, Phytopythium iriomotense sp. nov. and P. aichiense sp. nov., isolated from river water and water purification sludge in Japan. Mycological Progress 14: 1–12. Bezerra JDP, Sandoval-Denis M, Paiva LM, et al. 2017. New endophytic Toxicocladosporium species from cacti in Brazil, and description of Neocladosporium gen. nov. IMA Fungus 8: 77– 97. Bon M. 1984. Novitates – Combinaisons et taxons nouveaux. Documents Mycologiques 14 (53): 6. Borchsenius F. 2009. FastGap 1.2. Department of Bio-sciences, Aarhus University, Denmark. http://www.aubot.dk/FastGap_home.htm. Brasier CM, Sanchez-Hernandez E, Kirk SA. 2003. Phytophthora inundata sp. nov., a part heterothallic pathogen of trees and shrubs in wet or flooded soils. Mycological Research 107: 477– 484. Braun U, Crous PW, Nakashima C. 2015. Cercosporoid fungi (Mycosphaerellaceae) 3. Species on monocots (Poaceae, true grasses). IMA Fungus 6: 25 – 97. Braun U, Hill CF. 2002. Some new micromycetes from New Zealand. Mycological Progress 1: 19 – 30. Burgess TI, Simamora A, White D, et al. 2018. New species from Phytophthora Clade 6a: evidence for recent radiation. Persoonia 41: 1–17. Buyck B. 2000. Le genre Cantharellus en Europe. Bulletin de la Société Mycologique de France 116: 91–1378. Buyck B, Hofstetter V. 2011. The contribution of tef-1 sequences for species delimitation in the Cantharellus cibarius complex in the southeastern USA. Fungal Diversity 49: 35 – 46. Ceruti A, Fontana A, Nosenzo C. 2003. Le specie europee del genere Tuber. Una revision storica. Museo Regionale di Scienze Naturali, Torino. Chaverri P, Liu M, Hodge KT. 2008. A monograph of the entomopathogenic genera Hypocrella, Moelleriella, and Samuelsia gen. nov. (Ascomycota, Hypocreales, Clavicipitaceae), and their aschersonia-like anamorphs in the Neotropics. Studies in Mycology 60: 1– 66. Chevenet F, Brun C, Bañuls A-L, et al. 2006. TreeDyn: towards dynamic graphics and annotations for analyses of trees. BMC Bioinformatics 7: 439. Clémençon H. 2012. Cytology and plectology of the hymenomycetes. 2nd ed. Cramer, Stuttgart. Corner EHJ. 1994. Agarics in Malesia II Mycenoid. Beihefte zur Nova Hedwigia 109: 1– 271. Crous PW. 2002. Taxonomy and pathology of Cylindrocladium (Calonectria) and allied genera. APS Press. Crous PW, Braun U, Schubert K, et al. 2007a. Delimiting Cladosporium from morphologically similar genera. Studies in Mycology 58: 33 – 56. Crous PW, Decock C, Schoch CL. 2001. Xenocylindrocladium guianense and X. subverticillatum, two new species of hyphomycetes from plant debris in the tropics. Mycoscience 42: 559 – 566. Crous PW, Groenewald JZ. 2016. They seldom occur alone. Fungal Biology 120: 1392 –1415. Crous PW, Groenewald JZ. 2017. The genera of fungi – G 4: Camarosporium and Dothiora. IMA Fungus 8: 131–152. Crous PW, Groenewald JZ, Shivas RG. 2011. Fungal Planet description sheets: 69 – 91. Persoonia 26: 108 –156. Crous PW, Luangsa-ard JJ, Wingfield MJ, et al. 2018a. Fungal Planet description sheets: 785 – 867. Persoonia 41: 238 – 417. Crous PW, Mohammed C, Glen M, et al. 2007b. Eucalyptus microfungi known from culture. 3. Eucasphaeria and Sympoventuria genera nova, and new species of Furcaspora, Harknessia, Heteroconium and Phacidiella. Fungal Diversity 25: 19 – 36. Crous PW, Schumacher RK, Akulov A, et al. 2019a. New and interesting fungi. 2. Fungal Systematics and Evolution 3: 57–134. Crous PW, Schumacher RK, Wingfield MJ, et al. 2018b. New and interesting fungi. 1. Fungal Systematics and Evolution 1: 169 – 215. Crous PW, Shivas RG, Quaedvlieg W, et al. 2014. Fungal Planet description Sheets: 214 – 280. Persoonia 32: 184 – 306. Crous PW, Shivas RG, Wingfield MJ, et al. 2012. Fungal Planet description sheets: 128 –153. Persoonia 29: 146 – 201. Crous PW, Wingfield MJ, Burgess TI, et al. 2017a. Fungal Planet description sheets: 625 –715. Persoonia 39: 270 – 467. Crous PW, Wingfield MJ, Burgess TI, et al. 2018c. Fungal Planet description sheets: 716 –784. Persoonia 40: 240 – 393. Crous PW, Wingfield MJ, Burgess TI. 2017b. Fungal Planet description sheets: 558 – 624. Persoonia 38: 240 – 384. Crous PW, Wingfield MJ, Cheewangkoon R, et al. 2019b. Foliar pathogens of eucalypts. Studies in Mycology 94: 125 – 298. Crous PW, Wingfield MJ, Chooi Y-H, et al. 2020a. Fungal Planet description sheets: 1042 –1111. Persoonia 44: 301– 459. Crous PW, Wingfield MJ, Lombard L, et al. 2019c. Fungal Planet description sheets: 951–1041. Persoonia 43: 223 – 425. Crous PW, Wingfield MJ, Schumacher RK, et al. 2020b. New and interesting fungi. 3. Fungal Systematics and Evolution 6: 157– 231. Da Cruz ACR, Gusmão LFP, Castañeda-Ruíz RF. 2007. Conidial fungi from the semi-arid Caatinga biome of Brazil. Subramaniomyces pulcher sp. nov. and notes on Sporidesmium circinophorum. Mycotaxon 102: 25 – 32. Daranagama DA, Camporesi E, Liu XZ, et al. 2016. Tristratiperidium microsporum gen. et sp. nov. (Xylariales) on dead leaves of Arundo plinii. Mycological Progress 15: 8. Darriba D, Taaboada GL, Doallo R, et al. 2012. jModelTest 2: more models, new heuristics and parallel computing. Nature Methods 9: 772. Das K, Atri NS, Buyck B. 2013. Three new species of Russula (Russulales) from Sikkim (India). Mycosphere 4: 722 –732. Das K, Rossi W, Leonardi M, et al. 2018. Fungal Biodiversity Profiles 61–70. Cryptogamie, Mycologie 39: 381– 418. Day MJ, Gibas CFC, Fujimura KE, et al. 2006. Monodictys arctica, a new hyphomycete from the roots of Saxifraga oppositifolia collected in the Canadian High Arctic. Mycotaxon 98: 261– 272. De Beer ZW, Duong TA, Barnes I, et al. 2014. Redefining Ceratocystis and allied genera. Studies in Mycology 79: 187– 219. De Hoog GS, Vicente VA, Najafzadeh MJ, et al. 2011. Waterborne Exophiala species causing disease in cold-blooded animals. Persoonia 27: 46 –72. De Hoog GS, Weenink XO, Gerrits van den Ende AHG. 1999. Taxonomy of the Phialophora verrucosa complex with the description of two new species. Studies in Mycology 43: 107–122. DeCock C, Hennebert GL, Crous PW. 1997. Nectria serpens sp. nov. and its hyphomycetous anamorph Xenocylindrocladium gen. nov. Mycological Research 101: 786 –790. Deighton FC. 1967. Studies on Cercospora and allied genera. II. Passalora, Cercosporidium and some species of Fusicladium on Euphorbia. Mycological Papers 112: 1– 80. Dennis RWG. 1952. Lepiota and allied genera in Trinidad, British West Indies. Kew Bulletin 7: 459 – 500. Dereeper A, Guignon V, Blanc G, et al. 2008. Phylogeny.fr: robust phylogenetic analysis for the non-specialist. Nucleic Acids Research 36 (Web Server issue), W465 ‒W469. Durand EJ. 1908. The Geoglossaceae of North America. Annales Mycologici 6: 387– 477. Edgar RC. 2004. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research 32: 1792 ‒1797. Fries EM. 1836–1838. Epicrisis systematis mycologici seu synopsis Hymenomycetum. Uppsala, Sweden. Gams W. 1971. Cephalosporium-artige Schimmelpilze (Hyphomycetes). Fischer, Stuttgart. Gams W, Holubová-Jechová V. 1976. Chloridium and some other dematiaceous hyphomycetes growing on decaying wood. Studies in Mycology 13: 1– 99. Garnica S, Weiß M, Oertel B, et al. 2009. Phylogenetic relationships in Cortinarius, section Calochroi, inferred from nuclear DNA sequences. BMC Evolutionary Biology 9: 1. Garzon LM, Rueda LJ, Celis AM, et al. 2019. Exophiala psychrophila: a new agent of chromoblastomycosis. Medical Mycology Case Reports 23: 31–33. Gelardi M, Simonini G, Ercole E, et al. 2014a. Alessioporus and Pulchroboletus (Boletaceae, Boletineae), two novel genera for Xerocomus ichnusanus and X. roseoalbidus from the European Mediterranean basin: molecular and morphological evidence. Mycologia 106: 1168 –1187. Fungal Planet description sheets Gelardi M, Vizzini A, Ercole E, et al. 2014b. New collection, iconography and molecular evidence for Tylopilus neofelleus (Boletaceae, Boletoideae) from southwestern China and the taxonomic status of T. plumbeoviolaceoides and T. microsporus. Mycoscience 56: 373 – 386. Gomes RR, Vicente VA, De Azevedo CMPS, et al. 2016. Molecular epidemiology of agents of human chromoblastomycosis in Brazil with the description of two novel species. PLoS Neglected Tropical Diseases 10 (11): e0005102. Guarro J, Gene J, Stchigel AM, et al. 2012. Atlas of soil ascomycetes. CBS Biodiversity Series 10. Westerdijk Fungal Biodiversity Institute, Utrecht, The Netherlands. Guevara-Suarez M, García D, Cano-Lira JF, et al. 2020. Species diversity in Penicillium and Talaromyces from herbivore dung, and the proposal of two new genera of penicillium-like fungi in Aspergillaceae. Fungal Systematics and Evolution 5: 39 –75. Guindon S, Dufayard JF, Lefort V, et al. 2010a. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Systematic Biology 59: 307– 321. Guindon S, Lethiec F, Duroux P, et al. 2010b. PHYML online – a web server for fast maximum likelihood-based phylogenetic inference. Nucleic Acids Research 33 (Web Server issue): W557–W559. Hakelier N. 1967. Three new Swedish species of Geoglossum. Svensk Botanisk Tidskrift 61: 419 – 424. Heilmann-Clausen J, Verbeken A, Vesterholt J. 1998. The genus Lactarius. Fungi of Northern Europe Vol 2. Hernández-Restrepo M, Gené J, Castañeda-Ruiz RF, et al. 2017. Phylogeny of saprobic microfungi from Southern Europe. Studies in Mycology 86: 53 – 97. Herrera M, Bandala VL, Montoya L. 2018. Cantharellus violaceovinosus, a new species from tropical Quercus forests in eastern Mexico. MycoKeys 32: 91–109. Hesseltine CW, Fennell DI. 1955. The genus Circinella. Mycologia 47: 193 – 212. Hongo T. 1966. Notulae Mycologicae (5). Memoirs of the Faculty of Education, Shiga University. Natural Science (Otsu) 16: 57– 62. Hongo T. 1973. Enumeration of the Hygrophoraceae, Boletaceae and Strobilomycetaceae. Mycological reports from New Guinea and the Solomon Islands (16 – 21). Bulletin of the National Science Museum (Tokyo) 16: 537– 557. Hosoya T, Huhtinen S. 2002. Hyaloscyphaceae in Japan (7): Hyaloscypha albohyalina var. monodictys var. nov. Mycoscience 43: 405 – 409. Houbraken J, Kocsubé S, Visagie CM, et al. 2020. Classification of Aspergillus, Penicillium, Talaromyces and related genera (Eurotiales): an overview of families, genera, subgenera, sections, series and species. Studies in Mycology 95: 5 –169. Huelsenbeck JP, Ronquist F. 2001. MrBayes: Bayesian inference of phylogenetic trees. Bioinformatics 17: 754 –755. Jaklitsch WM, Voglmayr H. 2012. Phylogenetic relationships of five genera of Xylariales and Rosasphaeria gen. nov. (Hypocreales). Fungal Diversity 52: 75 – 98. Johnson M, Zaretskaya I, Raytselis Y, et al. 2008. NCBI BLAST: a better web interface. Nucleic Acids Research 36 (suppl_2): W5 –W9. Johnston PR, Quijada L, Smith CA, et al. 2019. A multigene phylogeny toward a new phylogenetic classification of Leotiomycetes. IMA Fungus 10: 1. Katoh K, Rozewicki J, Yamada KD. 2019. MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Briefings in Bioinformatics 20: 1160 –1166. Katoh K, Standley DM. 2013. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Molecular Biology and Evolution 30: 772 –780. Kearse M, Moir R, Wilson A, et al. 2012. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28: 1647–1649. Kerekes JF, Desjardin DE. 2009. A monograph of the genera Crinipellis and Moniliophthora from Southeast Asia including a molecular phylogeny of the nrITS region. Fungal Diversity 37: 101‒152. Kirk PM. 1982. New or interesting microfungi IV. Dematiaceous hyphomycetes from Devon. Transactions of the British Mycological Society 78: 55–74. Kirk PM. 1983. New or interesting microfungi IX. Dematiaceous hyphomycetes from Esher Common. Transactions of the British Mycological Society 80: 449 – 467. Klaubauf S, Tharreau D, Fournier E, et al. 2014. Resolving the polyphyletic nature of Pyricularia (Pyriculariaceae). Studies in Mycology 79: 85 –120. Ko WH, Ann PJ. 1985. Phytophthora humicola, a new species from soil of a citrus orchard in Taiwan. Mycologia 77: 631– 636. Kornerup A, Wanscher JH. 1967. Methuen Handbook of Colour. Methuen & Co Ltd, London, England. Kornerup A, Wanscher JH. 1978. Methuen Handbook of Colour, 3rd ed. Eyre Methuen, London. © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 407 Kornerup A, Wanscher JH. 1983. Methuen Handbook of Colour, 3rd ed. Eyre Methuen, Ltd., London. Kumar S, Stecher G, Li M, et al. 2018. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Molecular Biology and Evolution 35: 1547–1549. Kumar S, Stecher G, Tamura K. 2016. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Molecular Biology and Evolution 33: 1870 –1874. Kurtzman CP. 2001. Six new anamorphic ascomycetous yeasts near Candida tanzawaensis. FEMS Yeast Research 1: 177–185. Kurtzman CP, Robnett CJ, Blackwell M. 2016. Description of Teunomyces gen. nov. for the Candida kruisii clade, Suhomyces gen. nov. for the Candida tanzawaensis clade and Suhomyces kilbournensis sp. nov. FEMS Yeast Research 16: 1– 9. Kušan I, Matočec N, Antonić O, et al. 2014. Biogeographical variability and re-description of an imperfectly known species Hamatocanthoscypha rotundispora (Helotiales, Hyaloscyphaceae). Phytotaxa 170: 1–12. Kušan I, Matočec N, Mešić A, et al. 2015. A new species of Thecotheus from Croatia with a key to the known species with apiculate spores. Sydowia 67: 51– 63. Lanfear R, Frandsen PB, Wright AM, et al. 2016. PartitionFinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Molecular Biology and Evolution 4: 772 –773. Leigh JW, Bryant D. 2015. POPART: full-feature software for haplotype network construction. Methods in Ecology and Evolution 6: 1110 –1116. Li GJ, Hyde KD, Zhao RL, et al. 2016. Fungal diversity notes 253 – 366: taxonomic and phylogenetic contributions to fungal taxa. Fungal Diversity 78: 1– 237. Li GJ, Zhao Q, Zhao D, et al. 2013. Russula atroaeruginea and R. sichuanensis spp. nov.from southwest China. Mycotaxon 124: 173 –188. Liu YJ, Whelen S, Hall BD, 1999. Phylogenetic relationships among ascomycetes: evidence from an RNA polymerase II subunit. Molecular Biology and Evolution 16: 1799e1808. Lombard L, Houbraken J, Decock C, et al. 2016. Generic hyper-diversity in Stachybotriaceae. Persoonia 36: 156 – 246. Lombard L, Van der Merwe NA, Groenewald JZ, et al. 2015. Generic concepts in Nectriaceae. Studies in Mycology 80: 189 – 245. Lorch JM, Lindner DL, Gargas A, et al. 2013. A culture-based survey of fungi in soil from bat hibernacula in the eastern United States and its implications for detection of Geomyces destructans, the causal agent of bat white nose syndrome. Mycologia 105: 237– 252. Maas Geesteranus RA, Horak E. 1995. Mycena and related genera from Papua New Guinea and New Caledonia. Bibliotheca Mycologica 159: 143 – 329. Maciá-Vicente JG, Glynou K, Piepenbring M. 2016. A new species of Exophiala associated with roots. Mycological Progress 15: 18. Madrid H, Hernández-Restrepo M, Gené J, et al. 2016. New and interesting chaetothyrialean fungi from Spain. Mycological Progress 15: 1179 –1201. Mains EB. 1957. Species of Cordyceps parasitic on Elaphomyces. Bulletin of the Torrey Botanical Club 84: 243 – 251. Man in’t Veld WA, Rosendahl KC, Brouwer H, et al. 2011. Phytophthora gemini sp. nov., a new species isolated from the halophilic plant Zostera marina in the Netherlands. Fungal Biology 115: 724 –732. Man in’t Veld WA, Rosendahl KC, Van Rijswick PC, et al. 2019. Multiple Halophytophthora spp. and Phytophthora spp. including P. gemini, P. inundata and P. chesapeakensis sp. nov. isolated from the seagrass Zostera marina in the Northern hemisphere. European Journal of Plant Pathology 153: 341– 57. Marin-Felix Y, Groenewald JZ, Cai L, et al. 2017a. Genera of phytopathogenic fungi: GOPHY 1. Studies in Mycology 86: 99 –216. Marin-Felix Y, Hernández-Restrepo M, Crous PW. 2020. Multi-locus phylogeny of the genus Curvularia and description of ten new species. Mycological Progress 19: 559 – 588. Marin-Felix Y, Senwanna C, Cheewangkoon R, et al. 2017b. New species and records of Bipolaris and Curvularia from Thailand. Mycosphere 8: 1556 –1574. Marincowitz S, Crous PW, Groenewald JZ, et al. 2008. Microfungi occurring on Proteaceae in the fynbos. CBS Biodiversity Series 7: 1–166. Massimo NC, Nandi Devan MM, Arendt KR, et al. 2015. Fungal endophytes in aboveground tissues of desert plants: infrequent in culture, but highly diverse and distinctive symbionts. Microbial Ecology 70: 61–76. Matheny PB, Swenie RA. 2018. The Inocybe geophylla group in North America: a revision of the lilac species surrounding I. lilacina. Mycologia 110: 618 – 634. Matsushima T. 1971. Microfungi of the Solomon Islands and Papua-New Guinea. Published by the author, Kobe. 408 McLeod A, Botha WJ, Meitz JC, et al. 2009. Biodiversity of Pythium species in South African agricultural systems. Mycological Research 113: 933–951. Mello A, Vizzini A, Longato S, et al. 2000. Tuber borchii versus Tuber maculatum: neotype studies and DNA analyses. Mycologia 92: 326 – 331. Miller MA, Pfeiffer W, Schwartz T. 2010. Creating the CIPRES Science Gateway for inference of large phylogenetic trees, In: Proceedings of the Gateway Computing Environments Workshop (GCE), 14 Nov. 2010, New Orleans, LA: 1– 8. Miller MA, Schwartz T, Pickett BE, et al. 2015. A RESTful API for access to phylogenetic tools via the CIPRES science gateway. Evolutionary Bioinformatics 11: EBO-S21501. Miller OK Jr, Lodge DJ, Baroni TJ. 2000. New and interesting ectomycorrhizal fungi from Puerto Rico, Mona, and Guana Islands. Mycologia 92: 558–570. Minh BQ, Nguyen MAT, Von Haeseler A. 2013. Ultrafast approximation for phylogenetic bootstrap. Molecular Biology and Evolution 30: 1188 –1195. Minnis AM, Lindner DL. 2013. Phylogenetic evaluation of Geomyces and allies reveals no close relatives of Pseudogymnoascus destructans, comb. nov., in bat hibernacula of eastern North America. Fungal Mycology 117: 638–649. Minter DW, Cannon PF. 2015. Geoglossum cookeanum. IMI Descriptions of Fungi and Bacteria, No. 204, Sheet 2031. Molia A, Larsson E, Jeppson M, et al. 2020. Elaphomyces section Elaphomyces (Eurotiales, Ascomycota) – taxonomy and phylogeny of North European taxa, with the introduction of three new species. Fungal Systematics and Evolution 5: 283 – 300. Morgan-Jones G. 1975. Notes on Hyphomycetes. VIII. Lylea, a new genus. Mycotaxon 3: 129 –132. Moser M. 1961 ‘1960’. Die Gattung Phlegmacium (Schleimköpfe). Die Pilze Mitteleuropas, Band IV. Klinkhardt, Germany. Mouzouras R, Jones EBG. 1985. Monodictys pelagica, the anamorph of Nereiospora cristata (Halosphaeriaceae). Canadian Journal of Botany 63: 2444 – 2447. Müller E, Von Arx JA. 1962. Die Gattungen der didymosporen Pyrenomyceten. Beiträge zur Kryptogamenflora der Schweiz 11: 1– 922. Munsell. 1975. Munsell Soil Colour Charts. Macbeth, Baltimore. Munsell AH, Munsell Color (Firm). 2000. Munsell soil color charts. New Windsor, NY: Munsell Color. Nagy LG, Desjardin DE, Vágvölgyi C, et al. 2013. Phylogenetic analyses of Coprinopsis section Lanatuli and Atramentarii identify multiple species within morphologically defined taxa. Mycologia 105: 112 –124. Najafzadeh MJ, Vicente VA, Feng P, et al. 2018. Rapid identification of seven waterborne Exophiala species by RCA DNA padlock probes. Mycopathologia 183: 669 – 677. Nakase T, Itoh M, Takematsu A, et al. 1988. Candida tanzawaensis, a new species of yeast isolated from moss collected in Japan. Transactions of the Mycological Society of Japan 29: 331– 338. Nannfeldt JA. 1942. The Geoglossaceae of Sweden (with regard also to the surrounding countries). Arkiv för Botanik 30A: 1– 67. Nguyen LT, Schmidt HA, Von Haeseler A, et al. 2015. IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Molecular Biology and Evolution 32: 268 – 274. Nishikawa J, Nakashima C. 2020. Japanese species of Alternaria and their species boundaries based on host range. Fungal Systematics and Evolution 5: 197– 281. Oberwinkler F. 1964. Intrahymeniale Heterobasidiomyceten: Fruchtkörperlose Sebacina-Sippen und ihre systematische Stellung. Nova Hedwigia 7: 489 – 499. Okonechnikov K, Golosova O, Fursov M, et al. 2012. Unipro UGENE: A unified bioinformatics toolkit. Bioinformatics 28: 1166 –1167. Paden JW. 1971. Three new species of Eupenicillium from soil. Mycopathologia et Mycologia Applicata 43: 259 – 268. Palmer JM, Kubatova A, Novakova A, et al. 2014. Molecular characterization of a heterothallic mating system in Pseudogymnoascus destructans, the fungus causing white-nose syndrome of bats. G3: Genes, Genomes, Genetics 4: 1755 –1763. Paz A, Bellanger JM, Lavoise C, et al. 2017. The genus Elaphomyces (Ascomycota, Eurotiales): a ribosomal DNA-based phylogeny and revised systematics of European ‘deer truffles’. Persoonia 38: 197– 239. Peck CH. 1985. Gymnascella. Mycotaxon 24: 68 – 93. Pegler DN. 1983. Agaric flora of the Lesser Antilles. Kew Bulletin Additional Series 9: 1– 668. Petch T. 1906. Descriptions of new Ceylon fungi. Annals of the Royal Botanic Gardens Peradeniya 3: 1–10. Pitt J. 1980. The genus Penicillium and its teleomorphic states Eupenicillium and Talaromyces. Academic Press, London. Pordel A, Khodaparast SA, McKenzie EHC, et al. 2017. Two new species of Pseudopyricularia from Iran. Mycological Progress 16: 729 –736. Raithelhuber J. 1974. Hongos argentinos. Tomo I. Buenos Aires: Compañía Impresora Argentina. Persoonia – Volume 45, 2020 Rayner RW. 1970. A Mycological Colour Chart. Commonwealth Mycological Institute, Kew and British Mycological Society. Réblová M, Hernández-Restrepo M, Fournier J, et al. 2020. New insights into the systematics of Bactrodesmium and its allies and introducing new genera, species and morphological patterns in the Pleurotheciales and Savoryellales (Sordariomycetes). Studies in Mycology 95: 415 – 466. Réblová M, Untereiner WA, Réblová K. 2013. Novel Evolutionary lineages revealed in the Chaetothyriales (Fungi) based on multigene phylogenetic analyses and comparison of ITS secondary structure. PLoS One 8 (5): e63547. Rehner SA, Buckley E. 2005. A Beauveria phylogeny inferred from nuclear ITS and EF1-α sequences: evidence for cryptic diversification and links to Cordyceps teleomorphs. Mycologia 97: 84 – 98. Riess K, Oberwinkler F, Bauer R, et al. 2014. Communities of endophytic Sebacinales associated with roots of herbaceous plants in agricultural and grassland ecosystems are dominated by Serendipita herbamans sp. nov. PLoS ONE 9: E94676. Roberts P. 1993. Exidiopsis species from Devon, including the new segregate genera Ceratosebacina, Endoperplexa, Microsebacina, and Serendipita. Mycological Research 97: 467– 478. Ronquist F, Huelsenbeck JP. 2003. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19: 1572 –1574. Ronquist F, Teslenko M, Van der Mark P, et al. 2012. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Systematic Biolology 61: 539 – 542. Ryberg M, Nilsson RH, Kristiansson E, et al. 2008. Mining metadata from unidentified ITS sequences in GenBank: a case study in Inocybe (Basidiomycota). BMC Evolutionary Biology 8: 50. Saar G, Brandrud TE, Dima B, et al. 2014. Cortinarius untergattung Phlegmacium sektion Purpurascentes in Europa. Journal des JEC 16: 140–161. Samson RA. 1972. Notes on Pseudogymnoascus, Gymnoascus and related genera. Acta Botanica Neerlandica 21: 517– 527. Samson RA, Yilmaz N, Houbraken J, et al. 2011. Phylogeny and nomenclature of the genus Talaromyces and taxa accommodated in Penicillium subgenus Biverticillium. Studies in Mycology 70: 159 –183. Sánchez-García M, Cifuentes-Blanco J, Matheny PB. 2013. Revisión taxonómica del género Melanoleuca en México y descripción de especies nuevas. Revista Mexicana de Biodiversidad 84 (supplemento micologia): 111–127. Sandoval-Denis M, Gené J, Sutton DA, et al. 2016. Redefining Microascus, Scopulariopsis and allied genera. Persoonia: Molecular Phylogeny and Evolution of Fungi. 36: 1– 36. Sarnari M.1998. Monografia illustrate del genere Russula in Europa. Tromo Primo, Italy. Sarwar S, Aziz T, Hanif M, et al. 2019. Russula swatica: A new species of Russula based on molecular, light microscopy, and scanning electron microscopy analyses from Swat Valley of Khyber Pakhtunkhwa province of Pakistan. Microscopy Research and Technique 82: 1700 –1705. Saxena AS, Mukerji KG. 1970. Fungi of Delhi XV. Lophotrichus indicus sp. nov. Acta Botanica Neerlandica 19: 722 –726. Seifert K, Morgan-Jones G, Gams W, et al. 2011. The genera of Hyphomycetes. CBS Biodiversity Series no. 9: 1–997. Westerdijk Fungal Biodiversity Institute, Utrecht, Netherlands. Silvestro D, Michalak I. 2012. raxmlGUI: a graphical front-end for RAxML. Organisms Diversity and Evolution 12: 335 – 337. Singer R. 1942. A monographic study of the genera Crinipellis and Chaetocalathus. Lilloa 8: 441‒ 534. Singer R. 1955. Type Studies on Basidiomycetes. VIII. Sydowia 9: 367‒431. Singer R, Digilio APL. 1951. Pródromo de la Flora Agaricina Argentina. Lilloa 25: 5 – 461. Singtripop C, Hongsanan S, Li J, et al. 2016. Chaetothyrina mangiferae sp. nov., a new species of Chaetothyrina. Phytotaxa 255: 21– 33. Soop K, Dima B, Cooper JA, et al. 2019. A phylogenetic approach to a global supraspecific taxonomy of Cortinarius (Agaricales) with an emphasis on the southern mycota. Persoonia 42: 261– 290. Stamatakis A. 2014. RAxML Version 8: A tool for phylogenetic analysis and postanalysis of large phylogenies. Bioinformatics 30: 1312 –1313. Stamatakis A. 2015. Using RAxML to infer phylogenies. Current Protocols in Bioinformatics 51: 6.14.1– 6.14.14. Stecher G, Tamura K, Kumar S. 2020. Molecular Evolutionary Genetics Analysis (MEGA) for macOS. Molecular Biology and Evolution 37: 1237–1239. Stevenson JA. 1975. Fungi of Puerto Rico and the American Virgin Islands. Contribution of the Reed Herbarium 23: 1–743. Suh S-Q, McHugh JV, Blackwell M. 2004. Expansion of the Candida tanzawaensis yeast clade: 16 novel Candida species from basidiocarp-feeding beetles. International Journal of Systematic and Evolutionary Microbiology 54: 2409 – 2429. Fungal Planet description sheets Summerbell RC, Gueidan C, Guarro J, et al. 2018. The protean acremonium. A. sclerotigenum/egyptiacum: Revision, food contaminant, and human disease. Microorganisms 6: 88. Swofford DL. 2002. PAUP*: Phylogenetic Analysis Using Parsimony (*and other methods). Version 4.0b10. Sunderland, Massachusetts: Sinauer Associates. Swofford DL. 2003. PAUP* 4.0b10. Phylogenetic Analysis Using Parsimony (*and other methods). Version 4. Sinauer Associates, Sunderland, MA, USA. Tamura K, Stecher G, Peterson D, et al. 2013. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Molecular Biology and Evolution 30: 2725 – 2729. Tanaka K, Hirayama K, Yonezawa H, et al. 2015. Revision of the Massarineae (Pleosporales, Dothideomycetes). Studies in Mycology 82: 75 –136. Tanney JB, Douglas B, Seifert KA. 2016. Sexual and asexual states of some endophytic Phialocephala species of Picea. Mycologia 108: 255 – 280. Tanney JB, Seifert KA. 2020. Mollisiaceae: An overlooked lineage of diverse endophytes. Studies in Mycology 95: 293 – 380. Taylor JE, Crous PW. 2001. Morphological variation and cultural characteristics of Coniothyrium leucospermi associated with leaf spots of Proteaceae. Mycoscience 42: 265 – 271. Thorn RG, Kim JI, Lebeuf R, et al. 2017. The golden chanterelles of Newfoundland and Labrador: a new species, a new record for North America, and a lost species rediscovered. Botany 95: 547– 560. Trifinopoulos J, Nguyen L-T, Von Haeseler A, et al. 2016. W-IQ-TREE: a fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Research: 44 (W1): W232 – W235. Tsuneda A, Hambleton S, Currah RS. 2004. Morphology and phylogenetic placement of Endoconidioma, a new endoconidial genus from trembling aspen. Mycologia 96: 1128 –1135. Uljé CB. 2005. Coprinus. In: Noordeloos ME, Kuyper TW, Vellinga EC (eds), Flora Agaricina Neerlandica 6: 22 –109. Taylor & Francis, Boca Raton. Van der Plaats-Niterink AJ. 1981. Monograph of the genus Pythium. Studies in Mycology 21: 1– 242. Varghese KI, Rao VG. 1980 ‘1979’. Forest microfungi I. Subramaniomyces, a new genus of hyphomycetes. Kavaka 7: 83 – 85. Velmurugan S, Prasannakumar C, Manokaran S, et al. 2013. DNA barcodes for marine fungal identification and discovery. Fungal Ecology 6: 408–418. Videira SIR, Groenewald JZ, Nakashima C, et al. 2017. Mycosphaerellaceae – chaos or clarity? Studies in Mycology 87: 257– 421. Visagie CM, Houbraken J, Frisvad JC, et al. 2014. Identification and nomenclature of the genus Penicillium. Studies in Mycology 78: 343 – 371. Vu D, Groenewald M, De Vries M, et al. 2019. Large-scale generation and analysis of filamentous fungal DNA barcodes boosts coverage for kingdom fungi and reveals thresholds for fungal species and higher taxon delimitation. Studies in Mycology 92: 135 –154. © 2020 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute 409 Wanasinghe DN, Hyde KD, Jeewon R, et al. 2017. Phylogenetic revision of Camarosporium (Pleosporineae, Dothideomycetes) and allied genera. Studies in Mycology 87: 207– 256. Warcup JH, Talbot PHB. 1967. Perfect states of Rhizoctonias associated with orchids. New Phytologist 66: 631– 641. Watling R, Gregory NM. 1986. Observations on the boletes of the Cooloola Sandmass, Queensland and notes on their distribution in Australia. Proceedings of the Royal Society of Queensland 97: 97–128. Weir BS, Johnston PR, Damm U. 2012. The Colletotrichum gloeosporioides species complex. Studies in Mycology 73: 115 –180. Weiss M, Waller F, Zuccaro A, et al. 2016. Sebacinales – one thousand and one interactions with land plants. New Phytologist 211: 20 – 40. Wendt L, Sir EB, Kuhnert E, et al. 2018. Resurrection and emendation of the Hypoxylaceae, recognised from a multigene phylogeny of the Xylariales. Mycological Progress 17: 115 –154. White TJ, Bruns T, Lee S, et al. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, et al. (eds), PCR protocols: A guide to methods and applications: 315 – 322. Academic Press, Inc., New York. Whitehead MR, Catullo RA, Ruibal M, et al. 2017. Evaluating multilocus Bayesian species delimitation for discovery of cryptic mycorrhizal diversity. Fungal Ecology 26: 74 – 84. Wijayawardene NN, Hyde KD, Bhat DJ, et al. 2015. Additions to brown spored coelomycetous taxa in Massarinae, Pleosporales: introducing Phragmocamarosporium gen. nov. and Suttonomyces gen. nov. Cryptogamie, Mycologie 36: 213 – 224. Woudenberg JHC, Hanse B, Van Leeuwen GCM, et al. 2017. Stemphylium revisited. Studies in Mycology 87: 77–103. Woudenberg JHC, Seidl MF, Groenewald JZ, et al. 2015. Alternaria section Alternaria: Species, formae speciales or pathotypes? Studies in Mycology 82: 1– 21. Wu G, Li YC, Zhu XT, et al. 2016. One hundred noteworthy boletes from China. Fungal Diversity 81: 25 –188. Yilmaz N, Visagie CM, Houbraken J, et al. 2014. Polyphasic taxonomy of the genus Talaromyces. Studies in Mycology 78: 175 – 341. Zheng J, Lu YF, Liu XJ, et al. 2017. Cyberlindnera xishuangbannaensis f.a., sp. nov., a yeast isolated from rotting wood. International Journal of Systematic and Evolutionary Microbiology 67: 5051– 5055. Zwickl DJ. 2006. Genetic algorithm approaches for the phylogenetic analysis of large biological sequence datasets under the maximum likelihood criterion. PhD dissertation, The University of Texas at Austen.