Academia.eduAcademia.edu
Fungal Diversity (2019) 96:1–242 https://doi.org/10.1007/s13225-019-00429-2 (0123456789().,-volV)(0123456789(). ,- volV) Fungal diversity notes 1036–1150: taxonomic and phylogenetic contributions on genera and species of fungal taxa Kevin D. Hyde1,2,3,4,5,6,71 • Danushka S. Tennakoon1,2,3,6,7 • Rajesh Jeewon8 • D. Jayarama Bhat9,10 • Sajeewa S. N. Maharachchikumbura11 • Walter Rossi12 • Marco Leonardi12 • Hyang Burm Lee13 • Hye Yeon Mun14 • Jos Houbraken15 • Thuong T. T. Nguyen13 • Sun Jeong Jeon13 • Jens Christian Frisvad16 • Dhanushka N. Wanasinghe1,3,4,71 • Robert Lücking17 • André Aptroot18 • Marcela E. S. Cáceres19 • Samantha C. Karunarathna1,4,5,71 • Sinang Hongsanan3,20 • Rungtiwa Phookamsak1,3,4,5,71 • Nimali I. de Silva1,3,5 • Kasun M. Thambugala21 • Ruvishika S. Jayawardena3 • Indunil C. Senanayake3,20 • Saranyaphat Boonmee3 • Jie Chen22 • Zong-Long Luo23 • Chayanard Phukhamsakda2,3 • Olinto L. Pereira24 • Vanessa P. Abreu25 • André Wilson Campos Rosado24 • Buyck Bart26 • Emile Randrianjohany27 • Valérie Hofstetter28 • Tatiana B. Gibertoni29 • Adriene Mayra da Silva Soares30 • Helio Longoni Plautz Jr.31 • Helen Maria Pontes Sotão30 • William Kalhy Silva Xavier32 • Jadson Diogo Pereira Bezerra33 • Thays Gabrielle Lins de Oliveira33 • Cristina Maria de Souza-Motta33 • Oliane Maria Correia Magalhães33 • Digvijayini Bundhun3,34 • Dulanjalee Harishchandra2,3,35 • Ishara S. Manawasinghe2,3,35 • Wei Dong3,6,34,36 • Sheng-Nan Zhang3,34 • Dan-Feng Bao3,23,34 • Milan C. Samarakoon3,5,37 • Dhandevi Pem2,3,6,20 • Anuruddha Karunarathna1,3,7,34 • Chuan-Gen Lin2,3,6 • Jing Yang2,3,6,37 • Rekhani H. Perera2,3,6,37 • Vinit Kumar3,34 • Shi-Ke Huang1,2,3,6 • Monika C. Dayarathne1,2,3,6 • Anusha H. Ekanayaka1,2,3 • Subashini C. Jayasiri1,3 • Yuanpin Xiao2,3,6,38 • Sirinapa Konta1,2,3,6 • Tuula Niskanen39 • Kare Liimatainen39 • Yu-Cheng Dai40 • Xiao-Hong Ji40 • Xue-Mei Tian41 • Armin Mešić42 • Sanjay K. Singh43 • Kunthida Phutthacharoen2,3,6 • Lei Cai4 • Touny Sorvongxay3 • Vinodhini Thiyagaraja1,3,6,34 • Chada Norphanphoun2,3,6,7,38 • Napalai Chaiwan1,2,3,6 • Yong-Zhong Lu3,6,38 • Hong-Bo Jiang1,2,3,6 • Jin-Feng Zhang3,37 • Pranami D. Abeywickrama2,3,35 • Janith V. S. Aluthmuhandiram2,3,35 • Rashika S. Brahmanage2,3,35 • Ming Zeng1,2,3,6 • Thilini Chethana2,3,35 • Deping Wei1,3,34 • Martina Réblová45 • Jacques Fournier46 • Jana Nekvindová47 • Renan do Nascimento Barbosa48 • José Ewerton Felinto dos Santos33 • Neiva Tinti de Oliveira33 • Guo-Jie Li44 • Damien Ertz49,50 • Qiu-Ju Shang2,3,37 • Alan J. L. Phillips51 • Chang-Hsin Kuo7 • Erio Camporesi52,53,54 • Timur S. Bulgakov55 • Saisamorn Lumyong3,5,68,69 • E. B. Gareth Jones3,56 • Putarak Chomnunti2,3 • Eleni Gentekaki2,3 • Frank Bungartz57,58,59 • Xiang-Yu Zeng3,38 • Sally Fryar60 • Zdenko Tkalčec42 • Junmin Liang44 • Guangshuo Li44,61 • Ting-Chi Wen38,62 • Paras Nath Singh43 • Yusufjon Gafforov63,64,70 • Itthayakorn Promputtha5,72 • Erandi Yasanthika2,3 • Ishani D. Goonasekara1,2,3 • Rui-Lin Zhao44 • Qi Zhao1 • Paul M. Kirk65 • Jian-Kui Liu37,66 • JiYe Yan35 • Peter E. Mortimer1,71 • Jianchu Xu1,4,71 • Mingkwan Doilom1,3,4,5,67,71 Received: 28 March 2019 / Accepted: 16 May 2019 / Published online: 24 June 2019  School of Science 2019 Abstract This article is the tenth series of the Fungal Diversity Notes, where 114 taxa distributed in three phyla, ten classes, 30 orders and 53 families are described and illustrated. Taxa described in the present study include one new family (viz. Pseudoberkleasmiaceae in Dothideomycetes), five new genera (Caatingomyces, Cryptoschizotrema, Neoacladium, Paramassaria and Trochilispora) and 71 new species, (viz. Acrogenospora thailandica, Amniculicola aquatica, A. guttulata, Angustimassarina sylvatica, Blackwellomyces lateris, Boubovia gelatinosa, Buellia viridula, Caatingomyces brasiliensis, Calophoma humuli, Camarosporidiella mori, Canalisporium dehongense, Cantharellus brunneopallidus, C. griseotinctus, Castanediella meliponae, Coprinopsis psammophila, Cordyceps succavus, Cortinarius minusculus, C. subscotoides, Diaporthe italiana, D. rumicicola, Diatrypella delonicis, Dictyocheirospora aquadulcis, D. taiwanense, Digi& Mingkwan Doilom todesmium chiangmaiense, Distoseptispora dehongensis, j_hammochi@hotmail.com D. palmarum, Dothiorella styphnolobii, Ellisembia aurea, Extended author information available on the last page of the article 123 2 Fungal Diversity (2019) 96:1–242 Falciformispora aquatic, Fomitiporia carpinea, F. lagerstroemiae, Grammothele aurantiaca, G. micropora, Hermatomyces bauhiniae, Jahnula queenslandica, Kamalomyces mangrovei, Lecidella yunnanensis, Micarea squamulosa, Muriphaeosphaeria angustifoliae, Neoacladium indicum, Neodidymelliopsis sambuci, Neosetophoma miscanthi, N. salicis, Nodulosphaeria aquilegiae, N. thalictri, Paramassaria samaneae, Penicillium circulare, P. geumsanense, P. mali-pumilae, P. psychrotrophicum, P. wandoense, Phaeoisaria siamensis, Phaeopoacea asparagicola, Phaeosphaeria penniseti, Plectocarpon galapagoense, Porina sorediata, Pseudoberkleasmium chiangmaiense, Pyrenochaetopsis sinensis, Rhizophydium koreanum, Russula prasina, Sporoschisma chiangraiense, Stigmatomyces chamaemyiae, S. cocksii, S. papei, S. tschirnhausii, S. vikhrevii, Thysanorea uniseptata, Torula breviconidiophora, T. polyseptata, Trochilispora schefflerae and Vaginatispora palmae). Further, twelve new combinations (viz. Cryptoschizotrema cryptotrema, Prolixandromyces australi, P. elongatus, P. falcatus, P. longispinae, P. microveliae, P. neoalardi, P. polhemorum, P. protuberans, P. pseudoveliae, P. tenuistipitis and P. umbonatus), an epitype is chosen for Cantharellus goossensiae, a reference specimen for Acrogenospora sphaerocephala and new synonym Prolixandromyces are designated. Twenty-four new records on new hosts and new geographical distributions are also reported (i.e. Acrostalagmus annulatus, Cantharellus goossensiae, Coprinopsis villosa, Dothiorella plurivora, Dothiorella rhamni, Dothiorella symphoricarposicola, Dictyocheirospora rotunda, Fasciatispora arengae, Grammothele brasiliensis, Lasiodiplodia iraniensis, Lembosia xyliae, Morenoina palmicola, Murispora cicognanii, Neodidymelliopsis farokhinejadii, Neolinocarpon rachidis, Nothophoma quercina, Peroneutypa scoparia, Pestalotiopsis aggestorum, Pilidium concavum, Plagiostoma salicellum, Protofenestella ulmi, Sarocladium kiliense, Tetraploa nagasakiensis and Vaginatispora armatispora). Keywords 71 new taxa  Ascomycota  Basidiomycota  Dothideomycetes  Eurotiomycetes  Lecanoromycetes  Leotiomycetes  Pezizomycetes  Phylogeny  Taxonomy Table of contents Pleosporomycetidae C.L. Schoch, Spatafora, Crous & Shoemaker Ascomycota R.H. Whittaker Pleosporales Luttr. ex M.E. Barr Amniculicolaceae Y. Zhang ter, C.L. Schoch, J. Fourn., Crous & K.D. Hyde 1039. Amniculicola aquatica Z.L. Luo, K.D. Hyde & H.Y. Su, sp. nov. (contribution by Zong-Long Luo/Dhanushka Wanasinghe) 1040. Amniculicola guttulata Z.L. Luo, K.D. Hyde & H.Y. Su, sp. nov. (contribution by Zong-Long Luo/Dhanushka Wanasinghe) 1041. Murispora cicognanii Wanas., Camporesi, E.B.G. Jones & K.D. Hyde, Cryptog. Mycol. 36: 437 (2015), new record from freshwater habitat (contribution by ZongLong Luo/Dhanushka Wanasinghe) Arthoniomycetes O.E. Erikss. & Wink Arthoniomycetidae P.M. Kirk, P. Cannon, Minter & Stalpers Arthoniales Henssen ex D. Hawksw. & O.E. Erikss Lecanographaceae Ertz, Tehler, G. Thor & Frisch (contribution by Damien Ertz/Frank Bungartz) 1036. Plectocarpon galapagoense Ertz & Bungartz, sp. nov Dothideomycetes O.E. Erikss. & Winka Dothideomycetidae P.M. Kirk, P.F. Cannon, J.C. David & Stalpers Capnodiales Woron Teratosphaeriaceae Crous & U. Braun 1037. Caatingomyces T.G.L. Oliveira, C.M. Souza-Motta, O.M.C. Magalhães & J.D.P. Bezerra, gen. nov. (contribution by Jadson Bezerra/Thays Gabrielle Lins de Oliveira/ Cristina Maria de Souza-Motta/Oliane Maria Correia Magalhães) 1038. Caatingomyces brasiliensis T.G.L. Oliveira, C.M. Souza-Motta, O.M.C. Magalhães & J.D.P. Bezerra, sp. nov. (contribution by Jadson Bezerra/Thays Gabrielle Lins de Oliveira/Cristina Maria de Souza-Motta/Oliane Maria Correia Magalhães) 123 Amorosiaceae Thambug. & K.D. Hyde 1042. Angustimassarina sylvatica N.I. de Silva, Camporesi & K.D. Hyde, sp. nov. (contribution by Nimali I. de Silva/Kasun M. Thambugala/Ruvishika S. Jayawardena) Camarosporidiellaceae Wanas., Wijayaw., Crous & K.D. Hyde 1043. Camarosporidiella mori Phutthacharoen, T.S. Bulgakov and K.D. Hyde, sp. nov. (contribution by Kunthida Phutthacharoen/Dhanushka Wanasinghe) Fungal Diversity (2019) 96:1–242 Cucurbitariaceae G. Winter 1044. Protofenestella ulmi Jaklitsch & Voglmayr, Stud. Mycol. 90: 111 (2017), new record (contribution by Digvijayini Bundhun/Dhanushka Wanasinghe/Rajesh Jeewon) Dictyosporiaceae Boonmee & K.D. Hyde 1045. Dictyocheirospora aquadulcis Sorvongxay, S. Boonmee & K.D Hyde, sp. nov. (contribution by Touny sorvongxay/Saranyaphat Boonmee) 1046. Dictyocheirospora rotunda D’souza, J. Bhat & K.D. Hyde, Fungal Divers. 80: 465 (2016), new geographical record (contribution by Shi-Ke Huang/Saranyaphat Boonmee) 1047. Dictyocheirospora taiwanense Tennakoon, C.H. Kuo & K.D. Hyde, sp. nov. (contribution by Danushka S. Tennakoon/Saranyaphat Boonmee) 1048. Digitodesmium chiangmaiense Shang & K.D. Hyde, sp. nov. (contribution by Qiu-Ju Shang/Saranyaphat Boonmee) Didymellaceae Gruyter, Aveskamp & Verkley 1049. Calophoma humuli Thiyagaraja, Bulgakov & K.D. Hyde, sp. nov. (contribution by Vinodhini Thiyagaraja/ Dhanushka Wanasinghe) 1050. Neodidymelliopsis farokhinejadii Ahmadp. & Mehr.-Koushk., Sydowia 69:175 (2017), new host record (contribution by Dulanjalee Harishchandra/Dhanushka Wanasinghe) 1051. Neodidymelliopsis sambuci Manawas., Camporesi & K.D. Hyde, sp. nov. (contribution by Ishara Manawasinghe/Dhanushka Wanasinghe) 1052. Nothophoma quercina (Syd. & P. Syd.) Q. Chen & L. Cai, Stud. Mycol. 82: 213 (2015), new host record (contribution by Dulanjalee Harishchandra/Dhanushka Wanasinghe) Hermatomycetaceae Locq. 1053. Hermatomyces bauhiniae Phukhams., D.J. Bhat & K.D. Hyde, sp. nov. (contribution by Chayanard Phukhamsakda/D. Jayarama Bhat) Lophiostomataceae K. Hiray. & Kaz. Tanaka 1054. Vaginatispora palmae S.N. Zhang, J.K. Liu & K.D. Hyde, sp. nov. (contribution by Shengnan Zhang/Jian-Kui Liu/Kasun M. Thambugala) 1055. Vaginatispora armatispora (K.D. Hyde, Vrijmoed, Chinnaraj & E.B.G. Jones) Wanas., E.B.G. Jones & K.D. Hyde, Studies in Fungi 1: 62 (2016), new record (contribution by Dan-Feng Bao/Kasun M. Thambugala) Massariaceae Nitschke 1056. Paramassaria Samarak., & K.D. Hyde, gen. nov. (contribution by Milan C. Samarakoon/Dhanushka Wanasinghe) 3 1057. Paramassaria samaneae Samarak & K.D. Hyde, sp. nov. (contribution by Milan C. Samarakoon/Dhanushka Wanasinghe) Phaeosphaeriaceae M.E. Barr 1058. Muriphaeosphaeria angustifoliae D. Pem, Gafforov & K.D. Hyde, sp. nov. (contribution by Dhandevi Pem/ Rajesh Jeewon) 1059. Neosetophoma miscanthi A. Karunarathna, C.H. Kuo & K.D. Hyde, sp. nov. (contribution by Anuruddha Karunarathna/CH Kuo/Rungtiwa Phookamsak) 1060. Neosetophoma salicis Norph., Gafforov, T.C. Wen & K.D. Hyde, sp. nov. (contribution by Chada Norphanphoun/Y. Gafforov/Sinang Hongsanan) 1061. Nodulosphaeria aquilegiae Chaiwan, Camporesi & K.D. Hyde, sp. nov. (contribution by Napalai Chaiwan/ Dhanushka Wanasinghe/Ruvishika S. Jayawardena) 1062. Nodulosphaeria thalictri D. Pem, Camporesi & K.D. Hyde, sp. nov. (contribution by Dhandevi Pem/Rajesh Jeewon) 1063. Phaeopoacea asparagicola Phukhams., A. Akulov & K.D. Hyde, sp. nov. (contribution by Chayanard Phukhamsakda/A. Akulov) 1064. Phaeosphaeria penniseti A. Karunarathna, C.H. Kuo & K.D. Hyde, sp. nov. (contribution by Anuruddha Karunarathna/C.H Kuo/Rungtiwa Phookamsak) Pseudoberkleasmiaceae Phukhams. & K.D. Hyde 1065. Pseudoberkleasmiaceae Phukhams. & K.D. Hyde, fam. nov. (contribution by Chayanard Phukhamsakda) 1066. Pseudoberkleasmium chiangmaiense Y.Z. Lu & K.D. Hyde, sp. nov. (contribution by Yong-Zhong Lu) Pyrenochaetopsidaceae Valenzuela-Lopez, Crous, Cano, Guarro & Stchigel 1067. Pyrenochaetopsis sinensis G.S. Li, J. M. Liang & L. Cai, sp. nov. (contribution by Lei Cai/Junmin Liang/ Guangshuo Li) Tetraplosphaeriaceae Kaz. Tanaka & K. Hiray. 1068. Tetraploa nagasakiensis (Kaz. Tanaka & K. Hirayama) Kaz. Tanaka & K. Hirayama, Fungal Divers. 63: 253 (2013), new geographical record (contribution by Hong-Bo Jiang/Rungtiwa Phookamsak) Torulaceae Corda 1069. Torula breviconidiophora C.G. Lin & K.D. Hyde, sp. nov. (contribution by Chuan-Gen Lin/Jian-Kui Liu) 1070. Torula polyseptata C.G. Lin & K.D. Hyde, sp. nov. (contribution by Chuan-Gen Lin/Jian-Kui Liu) Trematosphaeriaceae K.D. Hyde, Y. Zhang ter, Suetrong & E.B.G. Jones 1071. Falciformispora aquatica D.F. Bao, K.D. Hyde & H.Y. Su, sp. nov. (contribution by Dan-Feng Bao/Zonglong Liu) 123 4 Minutisphaerales Raja, Oberlies, Shearer & A.N. Mill. Acrogenosporaceae Jayasiri & K.D. Hyde 1072. Acrogenospora sphaerocephala (Berk. & Broome) M.B. Ellis, Dematiaceous Hyphomycetes: 114 (1971), reference specimen (contribution by Jing Yang/Subhashini Chathumini/Jian-Kui Liu) 1073. Acrogenospora thailandica J. Yang & K.D. Hyde, sp. nov. (contribution by Jing Yang/Subhashini Chathumini/Jian-Kui Liu) Dothideomycetes orders incertae sedis Asterinales M.E. Barr ex D. Hawksw. & O.E. Erikss. Asterinaceae Hansf. (= Lembosiaceae Hosag.) 1074. Lembosia xyliae X.Y. Zeng, T.C. Wen & K.D. Hyde, Fungal Divers. (2015), new host record (contribution by Xiang-Yu Zeng/Sinang Hongsanan) 1075. Morenoina palmicola J. Fröhl., K.D. Hyde & Joanne E. Taylor, Fungal Divers. 3:89 (2000), new host record (contribution by Sirirnapa Konta/Sinang Hongsanan) Botryosphaeriales C.L. Schoch, Crous & Shoemaker Botryosphaeriaceae Theiss. & P. Syd. 1076. Dothiorella plurivora Abdollahz., Javadi & A.J.L. Phillips, Persoonia 32:9 (2014), new record (contribution by Shi-Ke Huang/Pranami D. Abeywickrama/Jian-Kui Liu/ A.J.L. Phillips) 1077. Dothiorella rhamni Wanas., Bulgakov, E.B.G. Jones & K.D. Hyde, Fungal Divers. 78: 253 (2016), new host record (contribution by Janith Vishwakirthi/Pranami D. Abeywickrama/Jian-Kui Liu/A.J.L. Phillips) 1078. Dothiorella styphnolobii Brahmanage, Bulgakov & K.D. Hyde, sp. nov. (contribution by Rashika Brahmanage/ Pranami D. Abeywickrama/Timur S. Bulgakov/Jian-Kui Liu/A.J.L. Phillips) 1079. Dothiorella symphoricarposicola W.J. Li, J.K. Liu & K. D. Hyde, Cyptogamie Mycol 35: 265 (2015), new host record (contribution by Thilini Chethana/Pranami D. Abeywickrama/Jian-Kui Liu/A.J.L. Phillips) 1080. Lasiodiplodia iraniensis Abdollahz., Zare & A.J.L. Phillips, Persoonia 25: 8 (2010), new host, geographical record (contribution by Hansika Perera/Vinith Kumar/JianKui Liu/A.J.L. Phillips) Jahnulales K.L. Pang, Abdel-Wahab, El-Shar., E.B.G. Jones & Sivichai Aliquandostipitaceae Inderb. 1081. Jahnula queenslandica Dayar., Fryar & K.D. Hyde, sp. nov. (contribution by Monika Dayarathne/Sally Fryar/ Kevin D. Hyde) Tubeufiales Boonmee, K.D Hyde Tubeufiaceae M.E. Barr 1082. Kamalomyces mangrovei Dayar., & K.D. Hyde, sp. nov. (contribution by Monika Dayarathne/Kevin D. Hyde) 123 Fungal Diversity (2019) 96:1–242 Eurotiomycetes O.E. Erikss. & Winka Chaetothyriomycetidae Doweld Chaetothyriales M.E. Barr Herpotrichiellaceae Munk 1083. Thysanorea uniseptata N.G. Liu & K.D. Hyde, sp. nov. (contribution by Ningguo Liu/Jian-Kui Liu) Eurotiomycetidae Geiser & Lutzoni Eurotiales G.W. Martin ex Benny & Kimbr. Aspergillaceae Link (contribution by Hyang Burn Lee, Thuong T.T. Nguyen, Hye Yeon Mun, Jos Houbraken/Jens Christian Frisvad/Paul M. Kirk) 1084. Penicillium circulare Hyang B. Lee, P.M. Kirk & T.T.T. Nguyen, sp. nov. 1085. Penicillium geumsanense Hyang B. Lee, P.M. Kirk & T.T.T. Nguyen, sp. nov. 1086. Penicillium mali-pumilae Hyang B. Lee, T.T.T. Nguyen & Houbraken, sp. nov. 1087. Penicillium psychrotrophicum Hyang B. Lee, H.Y. Mun, J.C. Frisvad & Houbraken, sp. nov. 1088. Penicillium wandoense Hyang B. Lee, P.M. Kirk & T.T.T. Nguyen, sp. nov. Laboulbeniomycetes Engl. Laboulbeniales Lindau Laboulbeniaceae G. Winter (contribution by Walter Rossi/Marco Leonardi) 1089. Prolixandromyces R.K. Benjamin 1970, Aliso 7: 174 = Monandromyces R. K. Benjamin 1999, Aliso 18: 72, syn. nov. 1090. Prolixandromyces australis (R.K. Benjamin) W. Rossi, comb. nov. 1091. Prolixandromyces elongatus (R.K. Benjamin) W. Rossi, comb. nov. 1092. Prolixandromyces falcatus (R.K. Benjamin) W. Rossi, comb. nov. 1093. Prolixandromyces longispinae (R.K. Benjamin) W. Rossi, comb. nov. 1094. Prolixandromyces microveliae (Thaxter) W. Rossi, comb. nov. 1095. Prolixandromyces neoalardi (R.K. Benjamin) W. Rossi, comb. nov. 1096. Prolixandromyces polhemorum (R.K. Benjamin) W. Rossi, comb. nov. 1197. Prolixandromyces protuberans (R.K. Benjamin) W. Rossi, comb. nov. 1198. Prolixandromyces pseudoveliae (R.K. Benjamin) W. Rossi, comb. nov. 1199. Prolixandromyces tenuistipitis (R.K. Benjamin) W. Rossi, comb. nov. 1100. Prolixandromyces umbonatus (R.K. Benjamin) W. Rossi, comb. nov. Fungal Diversity (2019) 96:1–242 1101. Stigmatomyces chamaemyiae W. Rossi & M. Leonardi sp. nov. 1102. Stigmatomyces cocksii W. Rossi & M. Leonardi, sp. nov. 1103. Stigmatomyces papei W. Rossi & M. Leonardi, sp. nov. 1104. Stigmatomyces tschirnhausii W. Rossi & M. Leonardi, sp. nov. 1105. Stigmatomyces vikhrevii W. Rossi & M. Leonardi, sp. nov. Lecanoromycetes O.E. Erikss. & Winka Lecanoromycetidae P.M. Kirk et al. ex Miadl. Caliciales Bessey Caliciaceae Chevall. 1106. Buellia viridula Ekanayaka, & K.D. Hyde, sp. nov. (contribution by Hasini Ekanayaka) Lecanorales Nannf. Lecanoraceae Körb. 1107. Lecidella yunnanensis Ekanayaka, & K.D. Hyde, sp. nov. (contribution by Hasini Ekanayaka) Pilocarpaceae Zahlbr. 1108. Micarea squamulosa Aptroot, Lücking & M. Cáceres, sp. nov. (contribution by André Aptroot/Robert Lücking/Marcela E. S. Cáceres) Ostropomycetidae V. Reeb, Lutzoni & Cl. Roux Ostropales Nannf. Porinaceae Rchb. 1109. Porina sorediata Aptroot, Lücking & M. Cáceres, sp. nov. (contribution by André Aptroot/Robert Lücking/ Marcela E.S. Cáceres) Graphidaceae Dumort. 1110. Cryptoschizotrema Aptroot, Lücking & M. Cáceres, gen. nov. (contribution by André Aptroot/Robert Lücking/ Marcela E.S. Cáceres) 1111. Cryptoschizotrema cryptotrema (Nyl.) Aptroot, Lücking & M. Cáceres, comb. nov. (contribution by André Aptroot/Robert Lücking/Marcela E.S. Cáceres) Leotiomycetes O.E. Erikss. & Winka Helotiales Nannf. Chaetomellaceae Baral, P.R. Johnst. & Rossman 1112. Pilidium concavum (Desm.) Höhn., Sber. Akad. Wiss. Wien, Math.-naturw. Kl., Abt. 1 124: 148 (1915), new geographical record (contribution by Milan Samarakoon/Jian-Kui Liu) Pezizomycetes O.E. Erikss. & Winka Pezizales J. Schröt. Ascodesmidaceae J. Schröt. 5 1113. Boubovia gelatinosa M. Zeng, Q. Zhao & K.D. Hyde, sp. nov. (contribution by Ming Zeng/Hasini Ekanayaka) Sordariomycetes O.E. Erikss. & Winka Diaporthomycetidae Senan., Maharachch. & K.D. Hyde Diaporthales Nannf. Diaporthaceae Höhn. ex Wehm. 1114. Diaporthe italiana Chethana, Camporesi & K. D. Hyde, sp. nov. (contribution by Thilini Chethna/Chinthani Senanayaka) 1115. Diaporthe rumicicola Manawas., Camporesi & K.D. Hyde, sp. nov. (contribution by Ishara Manawasinghe/ Chinthani Senanayaka) Gnomoniaceae G. Winter 1116. Plagiostoma salicellum (Fr.) Sogonov, Stud. Mycol. 62: 73 (2008), new geographical record (contribution by Digvijayini Bundhun/Chinthani Senanayaka/Rajesh Jeewon) Diaporthomycetidae, families incertae sedis Distoseptisporaceae K.D. Hyde & McKenzie 1117. Distoseptispora dehongensis W. Dong, H. Zhang & K.D. Hyde, sp. nov. (contribution by Wei Dong/Mingkwan Doilom) 1118. Distoseptispora palmarum S.N. Zhang, K.D. Hyde & J.K. Liu, sp. nov. (contribution by Shengnan Zhang/JianKui Liu) Hypocreomycetidae O.E. Erikss. & Winka Glomerellales Chadef. ex Réblová, W. Gams & Seifert Plectosphaerellaceae W. Gams, Summerb. & Zare 1119. Acrostalagmus annulatus (Berk. & M.A. Curtis) Seifert, Stud. Mycol. 68: 186 (2011), new geographical record (contribution by Subhashini Chathumini/Hansika Perera/Ruvishika S. Jayawardena) Hypocreales Lindau Cordycipitaceae Kreisel ex G.H. Sung, J.M. Sung, HywelJones & Spatafora 1120. Cordyceps succavus Y.P. Xiao, T.C. Wen & K.D. Hyde, sp. nov. (contribution by Yuanpin xiao/Ting-Chi Wen/Sinang Hongsanan) 1121. Blackwellomyces lateris Y.P. Xiao, T.C. Wen & K.D. Hyde, sp. nov. (contribution by Yuanpin xiao/TingChi Wen/Sinang Hongsanan) Hypocreales genera incertae sedis 1122. Sarocladium kiliense (Grütz) Summerb., Stud. Mycol. 68: 158 (2011), new record (contribution by Deping Wei/Dhanushka Wanasinghe) Pleurotheciales Réblová & Seifert Pleurotheciaceae Réblová & Seifert 123 6 Fungal Diversity (2019) 96:1–242 1123. Phaeoisaria siamensis Jayasiri & K.D. Hyde, sp. nov. (contribution by Subashini Chathumini) Savoryellales Boonyuen, Suetrong, Sivichai, K.L. Pang & E.B.G. Jones Savoryellaceae Jaklitsch & Réblová 1124. Canalisporium dehongense W. Dong, H. Zhang & K.D. Hyde, sp. nov. (contribution by Wei Dong/Mingkwan Doilom) Sordariomycetidae O.E. Erikss. & Winka Chaetosphaeriales Huhndorf, A.N. Mill. Fernández & F.A. Chaetosphaeriaceae Réblová, M.E. Barr & Samuels 1125. Sporoschisma chiangraiense N.G. Liu & K.D. Hyde, sp. nov. (contribution by Ningguo Liu/Jian-Kui Liu) 1126. Ellisembia aurea Réblová & J. Fourn., sp. nov. (contribution by Réblová Martina/Jacques Fournier/Jana Nekvindová) Linocarpaceae Konta & K.D. Hyde 1127. Neolinocarpon rachidis Konta & K.D. Hyde, Mycosphere 8: 1968 (2017), new record (contribution by Sirinapa Konta/Saranyaphat Boonmee) Xylariomycetidae O.E. Erikss & Winka Amphisphaeriales D. Hawksw. & O.E. Erikss. Amphisphaeriaceae G. Winter 1128. Trochilispora V.P. Abreu, A.W.C. Rosado & O.L. Pereira, gen. nov. (contribution by Olinto Pereira/Abreu VP/Rosado AWC) 1129. Trochilispora schefflerae V.P. Abreu, A.W.C Rosado & O.L. Pereira, sp. nov. (contribution by Olinto Pereira/Abreu VP/Rosado A.W.C) Sporocadaceae Corda. 1130. Pestalotiopsis aggestorum F. Liu & L. Cai, Nature Scientific Reports 7: 870 (2017), new geographical record (contribution by Nimali I. De silva/Sajeewa Maharachchikumbura) Xylariaceae Tul. & C. Tul. 1134. Fasciatispora arengae Konta & K. D. Hyde, Mycosphere 9: 725 (2018), new record (contribution by Sirinapa Konta/Saranyaphat Boonmee) Basidiomycota R.T. Moore Agaricomycetes Doweld Agaricomycetidae Parmasto Agaricales Underw. Cortinariaceae R. Heim ex Pouzar (contribution by Tuula Niskanen/Kare Liimatainen) 1135. Cortinarius minusculus Liimat. & Niskanen, sp. nov. 1136. Cortinarius subscotoides Niskanen & Liimat., sp. nov. Psathyrellaceae Vilgalys et al. 1137. Coprinopsis psammophila Mešić & Tkalčec, sp. nov. (contribution by Armin Mešić/Zdenko Tkalčec) 1138. Coprinopsis villosa L. Nagy, Desjardin, Vágvölgyi & Papp, Mycologia 105(1): 120 (2013), new geographical record (contribution by Armin Mešić/Zdenko Tkalčec) Cantharellales Gäum. Botryobasidiaceae Jülich (contribution by Sanjay K. Singh/Paras Nath Singh) 1139. Neoacladium P.N. Singh & S.K. Singh, gen. nov. 1140. Neoacladium indicum P.N. Singh & S.K. Singh, sp. nov. Hydnaceae Chevall (contribution by Bart Buyck/E. Randrianjohany/V. Hofstetter) 1141. Cantharellus goossensiae (Beeli) Heinem., Bull. Jard. Bot. Etat Brux. 28: 406 (1958), epitypification and new records 1142. Cantharellus brunneopallidus Buyck, Randrianjohany & V. Hofst., sp. nov. 1143. Cantharellus griseotinctus Buyck, Randrianjohany & V. Hofst., sp. nov. Xylariales Nannf. Castanediellaceae Hern.-Restr., Guarro & Crous 1131. Castanediella meliponae J.E. Felinto-Santos, R.N. Barbosa & N.T. Oliveira, sp. nov. (contribution by Renan do Nascimento Barbosa/José Ewerton Felinto dos Santos/ Neiva Tinti de Oliveira) Hymenochaetales Oberw. Hymenochaetaceae Donk (contribution by Yu-Cheng Dai/ Xiao-Hong Ji/Xue-Mei Tian) 1144. Fomitiporia carpinea X.H. Ji, X.M. Tian & Y.C. Dai, sp. nov. 1145. Fomitiporia lagerstroemiae X.H. Ji, X.M. Tian & Y.C. Dai, sp. nov. Diatrypaceae Nitschke 1132. Diatrypella delonicis R.H. Perera & K.D. Hyde, sp. nov. (contribution by Hansika Perera/Monika Dayarathna) 1133. Peroneutypa scoparia Carmarán & A.I. Romero, Fungal Divers. 23: 84 (2006), new record from mangrove habitat (contribution by Vinith Kumar/Monika Dayarathna) Polyporales Gäum. Grammotheleaceae Jülich (contribution by Tatiana B. Gibertoni/A.M.S. Soares/Helio Longoni Plautz Jr/Helen Maria Pontes Sotão/William Kalhy Silva Xavier) 1146. Grammothele aurantiaca A.M.S. Soares, sp. nov. 1147. Grammothele micropora A.M.S. Soares & W.K.S. Xavier, sp. nov. 123 Fungal Diversity (2019) 96:1–242 1148. Grammothele brasiliensis Ryvarden, Syn. Fung. 33: 38 (2015), new record Russulales Kreisel ex P.M. Kirk, P.F. Cannon & J.C. David Russulaceae Lotsy (contribution by Guo-Jie Li/Rui-Lin Zhao) 1149. Russula prasina G.J. Li & R.L. Zhao, sp. nov. Chytridiomycota Arx Rhizophydiomycetes Tedersoo, Koljalg Rhizophydiales Letcher Rhizophydiaceae Werderm. (contribution by Hyang Burn Lee/Sun Jeong Jeon/Thuong T.T. Nguyen) 1150. Rhizophydium koreanum Hyang B. Lee, S.J. Jeon, T.T.T. Nguyen, sp. nov. Introduction Fungi play vital roles in all ecosystems, as decomposers, epiphytes, endophytes, other symbionts of plants, as well as animal and plant pathogens (de Silva et al. 2016; Liu et al. 2017; Hyde et al. 2018a). They are hyper-diverse and heterogeneous group of organisms that represent a large and diverse component of microbial diversity (Hawksworth 2001; Hernández-Restrepo et al. 2017). Fungi show a great variation in morphology, reproduction, life cycles and modes of dispersal (Promputtha et al. 2007; Lofgren et al. 2018). They exhibit different lifestyles, including being endophytes, biotrophs, hemi-biotrophs, necrotrophs and saprotrophs (de Silva et al. 2016). Fungal communities are a major component of soil biomass; they accelerate rock weathering and decay dead plant material (Kendrick 2000; Finlay 2008). They are an essential part of the ecosystem because they play an important role in the decomposition of organic materials (Soares et al. 2017; Asplund et al. 2018). The diversity of fungi has been estimated to range between 2.2 to 3.8 million (Hawksworth and Lücking 2017) however, only 120,000 species are presently known. Hence, many species are still undescribed. One reason behind this is due to the fact that many common fungi are cosmopolitan, having a wider geographical distribution than plants and other organisms. The fungi of many continents besides Europe have been poorly studied and in many cases European names have been given to nonEuropean taxa without in depth analysis. Fungal inventories in many countries and in their infancy. This was shown in a recent study on the fungi of northern Thailand where up to 96% of species in a genus of mushrooms were found to be new to science (Hyde et al. 2018b). Despite all challenges, researchers have established reliable ways to resolve fungal taxa at the ordinal, familial and generic, as 7 well as species levels (Hyde et al. 2013, 2017a, b, 2018b; Ariyawansa et al. 2015; Wanasinghe et al. 2018a; Phookamsak et al. 2019). Major results include 16 phyla being accepted in the kingdom fungi (Tedersoo et al. 2018; Wijayawardene et al. 2018a), an outline of the Ascomycota provided by Wijayawardene et al. (2018a) and an outline of the basal fungi by Wijayawardene et al. (2018b). In addition, Fungal Diversity notes (Ariyawansa et al. 2015; Hyde et al. 2016, 2018a; Wanasinghe et al. 2018a; Phookamsak et al. 2019), Fungal Planet (Crous et al. 2014) and Mycosphere notes series (Hyde et al. 2017b; Thambugala et al. 2017b; Jayawardena et al. 2018) have provided outlets for introducing the numerous new species of fungi. More than 1000 entries in Fungal Diversity notes introducing new taxa, reference specimens, new data, and other taxonomic contributions have been published with phylogenetic analyses. This is the tenth in the series of Fungal Diversity notes with more than 100 entries mainly collected from China, Italy, Russia, Thailand, plus some other Asian and European countries. The outcome of the series provides a stable taxonomy and phylogeny that can provide definitive classifications for researchers who need to accurately identify fungi in their research work. Materials and methods Materials and methods follow the previous fungal diversity notes (Hyde et al. 2016; Tibpromma et al. 2017; Wanasinghe et al. 2018a; Phookamsak et al. 2019). Ascomycota R.H. Whittaker Notes: We follow the latest treatments and updated accounts of Ascomycota in Wijayawardene et al. (2017a, 2018a). Class Arthoniomycetes O.E. Erikss. & Wink Arthoniomycetidae P.M. Kirk, P. Cannon, Minter & Stalpers Arthoniales Henssen ex D. Hawksw. & O.E. Erikss. Notes: The order Arthoniales is one of the main lichenized groups of the Ascomycota and belongs to the class Arthoniomycetes O.E. Erikss. & Winka in the Pezizomycotina O.E. Erikss. & Winka. The order currently includes the families Andreiomycetaceae B.P. Hodk. & Lendemer, Arthoniaceae Rchb. ex Rchb., Chrysotrichaceae Zahlbr., Lecanographaceae Ertz et al., Opegraphaceae Körb. ex Stizenb., Roccellaceae Chevall. and Roccellographaceae, and about 1500 accepted species (Lücking et al. 2017). Most species of Arthoniales form lichen symbioses, mainly with trentepohlioid algae, but 285 species are obligately lichenicolous (non-lichenized) (Diederich et al. 2018) and a few are considered as being doubtfully or not lichenized. 123 8 Lecanographaceae Ertz et al. Notes: The family Lecanographaceae was recognized in Ertz and Tehler (2011) and a formal description was given in Frisch et al. (2014) based on the type genus Lecanographa Egea & Torrente, along with Alyxoria Ach. ex Gray, Phacographa Hafellner, Plectocarpon Fée and Zwackhia Körb. Some Opegrapha species (viz. O. brevis and O. celtidicola) were placed in this family with molecular evidence but will have to be combined in another genus because the type of Opegrapha (O. vulgata) belongs to the Opegraphaceae. Plectocarpon Fée Notes: The genus Plectocarpon Fée was introduced in 1825 as a replacement for Delisea Fée, a later homonym of Delisea Lamouroux (Rhodophyta). Plectocarpon was first established for a lichen (now Pseudocyphellaria glabra) because Fée did not realized that the ascomata were those of a lichenicolous fungus. The lichenicolous habit was recognized 159 years later and the name Plectocarpon resurrected for a genus of lichenicolous fungi (Hawksworth and Galloway 1984; Galloway 2006). A world monograph of the genus accepted 32 species, 15 newly described (Ertz et al. 2005), while a few other species were added in recent years. The genus was placed in the Lecanographaceae with molecular evidence using Plectocarpon lichenum (Sommerf.) D. Hawksw. and P. nephromeum (Norman) R. Sant. (Frisch et al. 2014), but the generic type still needs to be sequenced. Plectocarpon galapagoense Ertz & Bungartz, sp. nov. Index Fungorum number: IF556360; Facesoffungi number: FoF05977; Fig. 1 Etymology: Referring to the occurrence in the Galapagos Islands. Holotype: CDS 33415 Ascomata lichenicolous, infecting and developing inside the host ascomata and thallus of Sarcographa tricosa sensu lato, at first immersed, later bursting through the host ascomata and thallus, black, epruinose, star-shaped becoming rounded, 1–2 mm diam., not gall-inducing, not constricted at the base, not producing necrotic areas; surface of lichenicolous ascomata plane to slightly convex, distinctly lirellate-labyrinthiform with slit-like hymenial disc, epruinose, typically delimited by remnants of lirellae and thallus cortex from the host, in particular along the ascomatal margins of the lichenicolous fungus. Stroma c. 90–130 lm thick; sterile stromatic hyphae dark brown to carbonized above, 15–40 lm thick, poorly developed between the loculi, absent below, K–. Hymenium hyaline, not inspersed with oil droplets, 65–100 lm high; hymenial gel I ? directly reddish or I ? blue turning reddish, K/I ? blue; epihymenium hyaline to pale brown, K–, I ? blue turning reddish with some parts remaining blue. 123 Fungal Diversity (2019) 96:1–242 Subhymenium hyaline or pale brown, K–, I ? blue with some parts turning reddish, 10–20 lm high. Paraphysoids sparingly branched in the hymenium, richly branched in the epihymenium, 1.5 lm thick, with broadened tips, 2.5–3 lm. Asci subcylindrical to narrowly clavate, with a narrow ocular chamber, with a tiny K/I ? blue ring in mature asci, (4–)6(–8)-spored, 55–70 9 11–14 lm. Ascospores fusiform, (2–)3-septate, septation starts with one median septum, the second cell often slightly larger, hyaline, becoming dark brown granulose when overmature, (15–)16.5–20.0(–22) 9 (4.5–)4.5–5.5(–6) lm (n = 50), l/b ratio 3.5–4.0; perispore hyaline, 1–1.5(–2) lm thick. Conidiomata sphaerical, immersed between the lirellae: wall dark brown, K–, c. 4–10 lm thick; conidia bacilliform, 4–5 9 1 lm. Material examined: ECUADOR, Galapagos Islands, Pinta Island, just below the highest point of the island, 0 350 200 N, 90 450 1300 W, 615 m elev., forest of Zanthoxylum fagara with abundant ferns in the understory, lichenicolous on Sarcographa tricosa sensu lato growing on twigs and branches of Chiococca alba, 26 February 2007, F. Bungartz 5759 A (CDS 33415, holotype); F. Bungartz 5755 A (CDS 33411, paratype); F. Bungartz 5760 A (CDS 33416, paratype). Distribution and ecology: Described as new from Galapagos and probably endemic; this lichenicolous fungus seems to be weakly parasitic, as no necrotic areas are visible on the hosts ascomata and thallus of Sarcographa tricosa sensu lato However, as the lichenicolous fungus frequently occur on the acomata of the host, it should incur a significant damage and deprive the host of some of its reproductive structures. The host lichen belongs to S. trigosa sensu lato (identification by Robert Lücking): excipulum very thin, c. 8–25 lm laterally; hymenium not inspersed, of c. 100–125 lm high; hypothecium hyaline, c. 12–25 lm; ascospores 5-septate, with thick septa, hyaline becoming greyish brown, c. 20–28 9 7–8 lm. Notes: In the key to Plectocarpon provided by Ertz et al. (2005), the new species would key out close to Plectocarpon macaronesiae Diederich, Etayo & Sérusiaux, but that species is endemic to Macaronesia and differs from P. galapagoense by smaller ascomata (up to 1.1 mm diam.), with a shiny rough surface, K ? green intensifying sterile stromatic hyphae and a different host genus (Lobaria). Among the recently described Plectocarpon species, P. dirinariae Ertz & van den Boom is the most similar to P. galapagoense but it differs by strongly convex and entirely non-carbonized ascomata and a different host genus (Dirinaria) (Ertz & van den Boom 2012). Plectocarpon aequatoriale Etayo, described from Ecuador, has notably strongly convex ascomata, longer ascospores and a different host genus (Sticta) (Etayo 2017). Opegrapha plectocarpoidea Diederich is lichenicolous on Phaeographis and Fungal Diversity (2019) 96:1–242 Fig. 1 Plectocarpon galapagoense (Bungartz 5759A, holotype). a–d Black ascomata emerging from the lirellate ascomata and the thallus of the host lichen (= Sarcographa tricosa sensu lato, with paler, distinctly pruinose disc). a, b Overview of the infected host lichen. c, d Detail of lichenicolous ascomata. c The left part of the lichenicolous 9 ascoma has been removed for microscopical examination. d The lichenicolous ascoma can be seen on the left, the still intact part of an ascoma of the host on the right. e Section through an ascoma of Plectocarpon, in water. f Ascus in KOH. g Ascospores in KOH. Scale bars: a, b = 2.5 mm, c, d = 500 lm, e = 100 lm, f, g = 10 lm 123 10 reminds of a species of Plectocarpon (Aptroot et al. 1997). That species, known from high altitude in Papua New Guinea, differs from Plectocarpon galapagoense by indistinctly lirellate, almost rounded ascomata with 8-spored asci and a dark brown exciple that continues below the hymenium. Class Dothideomycetes O.E. Erikss. & Winka Notes: Dothideomycetes is the largest class of Ascomycota. Liu et al. (2017) provided an updated multilocus phylogeny of the class Dothideomycetes with molecular dating evidence to unravel the evolutionary relationships and strengthen the ranking among taxa. In this study, we follow the classifications in the studies of Hyde et al. (2013), Wijayawardene et al. (2014) and Liu et al. (2017). The subclasses of Dothideomycetes and their families in each order are listed in alphabetical order. Dothideomycetidae P.M. Kirk, P.F. Cannon, J.C. David & Stalpers Notes: Dothideomycetidae was introduced by Kirk et al. (2001) and later validated by Schoch et al. (2006). According to the present status, Dothideomycetidae comprises the orders Capnodiales, Dothideales and Myriangiales which lack pseudoparaphyses and periphyses (Hyde et al. 2013). Capnodiales Woron. Notes: See Chomnunti et al. (2011) and Hyde et al. (2013) for details. Teratosphaeriaceae Crous & U. Braun Notes: The family Teratosphaeriaceae (Capnodiales, Dothideomycetes) represents complex cryptic taxa that are characterized by globular ascomata with brown walls of textura angularis, located in a stroma of pseudoparenchymatous cells (Crous et al. 2007). Teratosphaeriaceae comprises 59 quite diverse genera (Wijayawardene et al. 2018a), accommodating endophytic, saprobic, phytopathogenic, and opportunistic species (Crous et al. 2009; Quaedvlieg et al. 2014). We introduce a new genus Caatingomyces within Teratosphaeriaceae based on phylogenetic analyses of ITS and LSU rDNA sequence data, morphology and ecology. In this paper we introduce a new monotypic genus. Caatingomyces T.G.L. Oliveira, C.M. Souza-Motta, O.M.C. Magalhães & J.D.P. Bezerra, gen. nov. MycoBank number: MB827888; Facesoffungi number: FoF05818 Etymology: Named after its occurrence in the Brazilian tropical dry forest, Caatinga. Endophytic in plants. Sexual morph Undetermined. Asexual morph Mycelium superficial or immersed, hyaline to medium brown, septate, branched, hyphae thin-walled, 123 Fungal Diversity (2019) 96:1–242 broad and darker when adhered to the conidiomata. Conidiomata pycnidial, superficial or immersed, isolated or clustered, globose to subovoid, medium brown when young and dark at maturity, pseudoparenchymatous, glabrose, thin-walled. Neck absent or poorly developed, conidia released by irregular rupture. Conidiogenous cells phialidic, slightly tapered and truncated at the apex with robust base, sometimes branched, hyaline to medium brown, formed from the internal cells of the pycnidial wall. Conidia hyaline, aseptate, ellipsoid, obtuse apex with truncated base, smooth, thin-walled. Type: Caatingomyces brasiliensis T.G.L. Oliveira, C.M. Souza-Motta, O.M.C. Magalhães & J.D.P. Bezerra. Caatingomyces brasiliensis T.G.L. Oliveira, C.M. SouzaMotta, O.M.C. Magalhães & J.D.P. Bezerra, sp. nov. MycoBank number: MB827889; Facesoffungi number: FoF05817; Fig. 2 Etymology: Named after its occurrence in Brazil. Holotype: URM 91831 Endophyte of branches of Poincianella pyramidalis. Sexual morph Undetermined. Asexual morph Mycelium superficial or immersed, hyaline to medium brown, septate, branched, hyphae thin-walled, broad and darker when linked to the conidiomata. Conidiomata pycnidial (28.6–) 31.2–52 (–67.6) 9 (23.4–) 26–31.2 (– 52) lm, superficial or immersed in the culture medium, isolated or clustered, globose to subovoid, medium brown when young and dark at maturity, pseudoparenchymatous, glabrous, thin-walled. Rostrate ostiole absent or poorly developed, conidia released by irregular rupture. Conidiogenous cells (3–) 3.5–5 (–6.5) 9 (3–) 4–5.5 lm, phialidic, slightly tapered and truncated at apex with robust base, sometimes branched, hyaline to medium brown, arising from the internal cells of the pycnidial wall. Conidia (3.2–) 4.2–5.3 9 2.1 lm, hyaline, aseptate, ellipsoid, apex obtuse with truncated base, smooth and thin-walled. Culture characteristics: Colonies reaching 3 to 3.5 cm in diameter after 15 days at 25 C in the dark. On OA spreading, erumpent, aerial mycelium sparse with smooth margins; surface grey-olivaceous to white, reverse black. On PDA spreading, erumpent, aerial mycelium sparse with smooth and uneven margins; surface yellowish-brown to pale brown with presence of yellowish exudate, reverse pale brown centre to black at the extremities. On MEA spreading, initially creamy, erumpent, surface yellowish-whitish with smooth margins and presence of exudate, reverse beige. Material examined: BRAZIL, Paraı́ba State, Santa Teresinha municipally, Tamanduá farm, 071.524S, 03723.518 W, as endophyte from branches of Poincianella pyramidalis (Fabaceae), May 2013, J.D.P. Bezerra (URM 91831, holotype), ex-type living culture (URM 7916); as endophyte from branches of Poincianella Fungal Diversity (2019) 96:1–242 11 Fig. 2 Caatingomyces brasiliensis (URM 91831, holotype). a Colony on MEA after 7 days. b Colony on PDA after 7 days. c Colony on AO after 7 days. d Detail of colony on PDA after 7 days. e Pycnidia. f Detail of pycnidia. g Detail of pycnidial wall. h–l Conidiogenous cells and conidia. m Conidia. Scale bars: e = 50 lm, f = 25 lm, g–m = 10 lm pyramidalis (Fabaceae), May 2013, J.D.P. Bezerra, living culture (URM 7917). GenBank numbers: ITS: MH929437, MH929438, LSU: MH929439, MH929440, RPB2: MH929441, MH929442, TEF1-a: MH929443, MH929444. Notes: Maximum likelihood analysis using the alignment of LSU rDNA sequences as presented by Quaedvlieg et al. (2014), demonstrated relationships of the new genus with Readeriella Syd. & P. Syd. in Teratosphaeriaceae (data not shown). In the phylogenetic analyses (Fig. 3) using combined LSU and ITS sequence data of Teratosphaeriaceae, the new genus Caatingomyces is related to Readeriella species, but was placed in a distinct clade with high support. The TEF1-a and RPB2 sequences of C. brasiliensis were not included in the combined analyses, but are available in GenBank for further verification of its relationships and placement in Teratosphaeriaceae and affiliated families. Caatingomyces brasiliensis mainly differs from Readeriella species that have a thick conidiomata wall, the presence of single and circular ostiole, by the shape of the conidiogenous cells producing three collarettes and deltoid, thick-walled and guttulate conidia with 3 lateral obtuse projections (Sutton 1980). Furthermore, Caatingomyces also differs from species of other genera in Teratosphaeriaceae, such as Parapenidiella Crous & Summerell which has conidia arranged in branched acropetal chains, pigmented ramoconidia, and erect and macronematous conidiophores (Crous et al. 2012); from Eupenidiella Quaedvl. & Crous by erect and solitary, terminally penicillate conidiophores and subcylindrical conidia arranged in branched chains and secondary conidia (Quaedvlieg et al. 2014). Subclass Pleosporomycetidae C.L. Schoch et al. Pleosporales Luttr. ex M.E. Barr Notes: Pleosporales was established by Luttrell (1955) to accommodate members of Dothideomycetes having perithecioid ascomata with pseudoparaphyses amongst the asci (Zhang et al. 2009a). Pleosporales members are highly diverse and can be endophytes or epiphytes, parasitices, lichenicolous or saprobes in terrestrial or aquatic 123 12 Fungal Diversity (2019) 96:1–242 Fig. 3 Phylogram generated from Bayesian inference analysis based on combined LSU and ITS sequence data from 45 representative members of Teratosphaeriaceae. Related sequences are taken from Quaedvlieg et al. (2014). Forty-eight strains are included in the combined gene analyses comprising 1326 characters after alignment (708 characters for LSU, 618 characters for ITS). Parastagonospora nodorum (CBS 110109) (Phaeosphaeriaceae, Pleosporales) is used as the outgroup taxon. Analyses of each single gene were performed and the topology of each tree had clade stability. The tree topology of the Bayesian analysis was similar to the maximum likelihood analysis. In the Bayesian inference analysis presented, the estimated base frequencies for the LSU region were as follows: A = 0.2445, C = 0.2356, G = 0.3014, T = 0.2186; substitution rates AC = 0.5614, AG = 1.2118, AT = 0.8832, CG = 0.6931, CT = 5.0186, GT = 1.0000; range distribution parameter a = 0.6982. The estimated base frequencies for the ITS region were as follows: A = 0.1855, C = 0.3132, G = 0.2366, T = 0.2647; substitution rates AC = 2.1196, AG = 3.1193, AT = 2.6297, CG = 2.0375, CT = 5.3598, GT = 1.0000; range distribution parameter a = 0.4064. The best ranking RaxML tree with a final probability value was of - 8679.44. The matrix presented 1063 distinct alignment patterns, with 35.42% of characters or indeterminate intervals. Branches with later Bayesian probabilities (BYPP) equal to or greater than 0.95 and bootstrap support values for maximum likelihood (ML) analysis greater than 70 are placed above or below nodes. Ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue environments or even occur on animal dung (Zhang et al. 2009a). Among the orders of Dothideomycetes, Pleosporales is the largest and most diverse order containing more than 75 families (Wijayawardene et al. 2018a). pseudoparaphyses and short-pedicellate asci, bearing hyaline, reddish-brown or pale, 1- to multi-septate or muriform ascospores, generally with a hyaline gelatinous sheath’’ (Zhang et al. 2008). Amniculicolaceae Y. Zhang ter, C.L. Schoch, J. Fourn., Crous & K.D. Hyde Notes: The family Amniculicolaceae was introduced by Zhang et al. (2009a) to describe various freshwater taxa from Europe and later accepted by Shearer et al. (2009) with a well-supported phylogeny. The family is characterized by ‘‘ascomata with a rough black surface, usually staining the woody substrate purple, narrow Amniculicola Y. Zhang ter & K.D. Hyde Notes: Amniculicola was introduced to accommodate A. lignicola, which was isolated from submerged wood in freshwater in France (Zhang et al. 2008). This genus is characterised by ascomata with slit-like ostioles, thin, branching and anastomosing pseudoparaphyses, cylindrical asci, and hyaline, 1–3-septate ascospores. Presently, four species, viz. Amniculicola immersa, A. lignicola, A. 123 Fungal Diversity (2019) 96:1–242 longissima, A. parva are accepted in Amniculicola and all species were collected from freshwater habitats. Amniculicola aquatica Z.L. Luo, K.D. Hyde & H.Y. Su, sp. nov. Index Fungorum number: IF555506; Facesoffungi number: FoF05308; Fig. 4 Etymology: Referring to the aquatic habitat of this fungus. Holotype: MFLU 18–1324 Saprobic on decaying wood submerged in freshwater. Sexual morph Ascomata 260–330 lm high, 250–320 lm diam., solitary, scattered, or in small groups, superficial, coriaceous, with basal wall remaining immersed in host tissue, globose to subglobose, black, often laterally flattened, with a flattened base not easily removed from the substrate, mostly bearing remnants of wood fibres. Peridium 35–50 lm thick, 2-layered, outer layer composed of small, dark brown to black, heavily pigmented, thickwalled cells of textura angularis, cells at apex smaller and walls thicker, inner layer composed of hyaline, thin-walled cells of textura angularis. Hamathecium of dense, very long trabeculate pseudoparaphyses. Asci 110–130(– 152) 9 10–13 lm ( x = 120 9 11.5 lm, n = 10), 8spored, bitunicate, fissitunicate, cylindrical, with a short, narrowed, twisted, bifurcate pedicel. Ascospores 24– 32 9 6–8 lm ( x = 28 9 7 lm, n = 25), obliquely uniseriate and partially overlapping, fusiform, with narrowly rounded to acute ends, hyaline, uniseptate, deeply constricted at the septum, sometimes curved, smooth, with 4–6 prominent guttules, and a wide gelatinous sheath. Asexual morph Undetermined. Material examined: CHINA, Yunnan Province, saprobic on decaying wood submerged in Erhai Lake, March 2015, Z.L. Luo, S-318 (MFLU 18-1324, holotype; HKAS 92579, isotype), ex-type living culture (MFLUCC 16-1123); saprobic on decaying wood submerged in Erhai Lake, March 2015, H.Y. Su, S-467 (MFLU 18-1325, paratype; HKAS 92559), ex-paratype living culture (MFLUCC 16-0915). GenBank numbers: LSU: MK106096, TEF1-a: MK109800 (MFLUCC 16-1123); LSU: MK106097, SSU: MK106108, TEF1-a: MK109801 (MFLUCC 16-0915). Notes: Amniculicola aquatica mostly resembles A. parva in having superficial, coriaceous ascomata of similar size, cylindrical asci with a short, narrowed, twisted, bifurcate pedicel and fusiform, hyaline ascospores with a wide gelatinous sheath. However, A. aquatica differs from A. parva has a thicker peridium, and uniseptate, wider ascospores (6–8 vs. 4.5–6.5 lm) (Zhang et al. 2009b). Phylogenetic analysis based on LSU, SSU and TEF1-a sequence data showed that A. aquatica is a distinct taxon in the genus Amniculicola (Fig. 7). 13 Amniculicola guttulata Z.L. Luo, K.D. Hyde & H.Y. Su, sp. nov. Index Fungorum number: IF555507; Facesoffungi number: FoF05309; Fig. 5 Etymology: Referring to the ascospores with large guttules. Holotype: MFLU 18–1326 Saprobic on decaying wood submerged in freshwater habitats. Sexual morph Ascomata 290–320 lm high, 280– 300 lm diam., solitary, scattered, or in small groups, superficial, coriaceous, with basal wall remaining immersed in host tissue, globose to subglobose, black, often laterally flattened, with a flattened base not easily removed from the substrate, mostly bearing remnants of wood fibres. Peridium 27–35 lm thick, 2-layered, outer layer composed of small, brown to dark brown, thickwalled cells of textura angularis, cells at apex smaller and walls thicker, inner layer composed of hyaline, thin-walled cells of textura angularis. Hamathecium of dense, very long trabeculate pseudoparaphyses. Asci 113–127 9 9– 11 lm ( x = 120 9 10 lm, n = 10), 8-spored, bitunicate, fissitunicate, cylindrical, with a short, narrowed, twisted, bifurcate pedicel. Ascospores 23–27 9 5–7 lm ( x = 25 9 6 lm, n = 25), obliquely uniseriate and partially overlapping, fusiform, with narrowly rounded to acute ends, hyaline, uniseptate, deeply constricted at the septum, sometimes curved, smooth, with 4–6 prominent guttules, and gelatinous sheath. Asexual morph Undetermined. Material examined: CHINA, Yunnan Province, saprobic on decaying wood submerged in Dulong River, May 2015, Z.L. Luo, S-428 (MFLU 18-1326, holotype; HKAS 92685, isotype), ex-type living culture (MFLUCC 16-0907); saprobic on decaying wood submerged in Dulong River, May 2015, Z.L. Luo, S-516, living culture MFLUCC 16-1297; saprobic on decaying wood submerged in Nujiang River, May 2015, Z.L. Luo, S-538 (MFLU 18-1327, paratype), ex-paratype living culture (DLUCC 0538). GenBank numbers: LSU: MK106098, TEF1-a: MK109802 (MFLUCC 16-0907); LSU MK106099 (MFLUCC 16-1297); LSU: MK106100, SSU: MK106109, TEF1-a: MK109803 (DLUCC 0538). Notes: The new species Amniculicola guttulata resembles A. aquatica in having superficial, coriaceous, globose to subglobose ascomata, cylindrical asci of similar size and with a short, narrowed, twisted, bifurcate pedicel, and fusiform, hyaline, uniseptate ascospores deeply constricted at the septum. However, A. guttulata differs from A. aquatica has a thinner peridium and smaller ascospores (23–27 9 5–7 vs. 24–32 9 6–8 lm). Phylogenetic analysis showed that isolates of A. guttulata formed a separate 123 14 Fungal Diversity (2019) 96:1–242 Fig. 4 Amniculicola aquatica (MFLU 18–1324, holotype). a Appearance of ascomata on host substrate. b Section through ascoma. c Structure of peridium. d Pseudoparaphyses. e, f Asci. g–k Ascospores. Scale bars: b = 150 lm, e, f = 40 lm, c, d = 20 lm, g–k = 15 lm clade in this genus with high support (98 ML/0.99 Bayesian). Murispora Y. Zhang bis, J. Fourn. & K.D. Hyde Notes: The genus Murispora was introduced by Zhang et al. (2009a) to accommodate Pleospora rubicunda Niessl which is characterized by immersed, erumpent or nearly superficial, globose to subglobose, elongate, weakly 123 papillate ascomata that stain the woody substrate purple, filamentous, narrow, branched, septate, pseudoparaphyses, 8-spored, bitunicate, cylindro-clavate asci, and oval to ellipsoidal or fusiform, pale or reddish brown, asymmetrical, muriform ascospores, with one side flattened. Wanasinghe et al. (2015) introduced six species for this genus which were collected from Italy and the UK. In this Fungal Diversity (2019) 96:1–242 15 Fig. 5 Amniculicola guttulata (MFLU 18–1326, holotype). a, b Appearance of ascomata on host substrate. c Section through ascoma. d, e Structure of peridium. f, g Asci. h–l Ascospores. Scale bars: c = 150 lm, f, g = 30 lm, d, e = 20 lm, h–l = 10 lm paper, we report on Murispora cicognanii as a new record for China. Murispora cicognanii Wanasinghe, Camporesi, E.B.G. Jones & K.D. Hyde, Cryptog. Mycol. 36: 437 (2015) Facesoffungi number: FoF01108; Fig. 6 Saprobic on decaying wood submerged in freshwater habitats. Sexual morph Ascomata 200–260 lm high, 160– 250 lm diam., solitary, globose to subglobose, dark brown to black, semi-immersed to immersed, substrate stained purple, fused to the host tissue. Peridium comprising 3–4 layers of brown to reddish brown cells of textura angularis, 123 16 Fungal Diversity (2019) 96:1–242 with inner 1–2 layers of cells thin-walled and hyaline. Hamathecium comprising numerous, filamentous, branched, septate, pseudoparaphyses. Asci 120–150 9 20– 25 lm ( x = 135 9 22.5 lm, n = 10), 8-spored, bitunicate, fissitunicate, cylindric-clavate, short pedicellate, thickwalled at the apex. Ascospores 32–36 9 11–15 lm ( x = 34 9 13 lm, n = 20), overlapping uniseriate, golden yellow turning brown when mature, fusiform, asymetrical with one side flattened, muriform with 5–9 transverse septa, and 1–2 longitudinal septa, slightly constricted at the middle septum, conical and narrowly rounded at the ends, surrounded by a mucilaginous sheath. Asexual morph Undetermined. Material examined: CHINA, Yunnan Province, saprobic on decaying wood submerged in a freshwater stream in Cangshan Mountain, May 2016, F. Yang, S-757, living culture (MFLUCC 17–1682). GenBank numbers: LSU: MK106101, SSU: MK106110, TEF1-a: MK109804. Notes: Murispora cicognanii was introduced by Wanasinghe et al. (2015) based on a collection from Italy and is only known from the type locality. This species is characterized by globose to subglobose, immersed ascomata, cylindric-clavate, short pedicellate asci with a minute ocular chamber and golden yellow, fusiform, asymmetrical, muriform ascospores, turning brown when mature, with one side flattened and surrounded by a mucilaginous Fig. 6 Murispora cicognanii (S-757, new record). a Appearance of ascomata on host substrate. b Section through ascoma. c Structure of peridium. d Pseudoparaphyses. e Asci. f–i Ascospores. j, k Colonies on MEA from surface and reverse. Scale bars: b = 100 lm, e = 50 lm, c, d = 20 lm, f–i = 15 lm 123 Fungal Diversity (2019) 96:1–242 sheath (Wanasinghe et al. 2015). The morphology of our new isolate fits well with M. cicognanii. Phylogenetic analysis of LSU, SSU and TEF1-a sequence data shows that our isolate clustered with M. cicognanii with good support (100 ML/1.00 BYPP) (Fig. 7). Based on morphology and phylogeny, we identify our isolate as M. cicognanii. This is first record for China and the first collection from freshwater. Amorosiaceae Thambug. & K.D. Hyde Notes: Thambugala et al. (2015) introduced Amorosiaceae to accommodate the genera Amorosia Mantle & D. Hawksw and Angustimassarina Thambuga., Kaz. Tanaka & K.D. Hyde. Amorosiaceae is characterized by immersed to semiimmersed ascomata with crest-like, papillate ostiole and hyaline, 1–3-septate ascospores with mucilaginous sheath and hyphomycetous asexual morph (Thambugala et al. 2015). Angustimassarina Thambugala, Kaz. Tanaka & K.D. Hyde Notes: Thambugala et al. (2015) introduced Angustimassarina based on the type species Angustimassarina populi Thambug. & K.D. Hyde. Most Angustimassarina species are fungicolous and they may be parasitic on other fungi and appear to grow within other ascomata of other ascomycetes (Hyde et al. 2017b). Currently, 10 epithets are listed in Index Fungorum (2019). In this study, we introduce a new Angustimassarina species; A. sylvatica on Fagus sylvatica from Italy. Angustimassarina sylvatica N.I. de Silva, Camporesi & K.D. Hyde, sp. nov. Index Fungorum number: IF556268; Facesoffungi number: FoF05992; Fig. 8 Etymology: The specific epithet reflects the host Fagus sylvatica. Holotype: MFLU 15-2603 Saprobic on dead twigs of Fagus sylvatica Sexual morph Ascomata 180–260 lm high, 150–200 lm diam. ( x = 200 9 180 lm, n = 5), uniloculate, scattered, immersed to semi-immersed in the plant host tissue, dark brown to black, globose to subglobose. Ostiole in the centre without a papilla. Peridium 8–12 lm wide, unequally thick, comprising brown cells of textura angularis of inner layer and 17–24 lm wide, unequally thick, comprising irregular dark brown cells of textura prismatica. Hamathecium comprising 1–1.5 lm septate, unbranched, cellular pseudoparaphyses, embedded in gelatinous matrix, between and above the asci. Asci 95–110 9 8– 12 lm ( x = 98 9 10 lm, n = 15), 8-spored, bitunicate, fissitunicate, cylindric-clavate, with short pedicel, rounded at the apex. Ascospores 21–25 9 4–5 lm ( x = 23 9 4 lm, n = 25), bi-seriate, hyaline, fusiform, 1– 17 septate with 2 pseudosepta, deeply constricted at the septum, widest at the centre and tapering toward the ends, straight, smooth-walled, guttulate, surrounded by a mucilaginous sheath. Asexual morph Undetermined. Material examined: ITALY, Province of Forlı̀-Cesena, Monte Fumaiolo - Verghereto, on dead land twigs of Fagus sylvatica (Fagaceae), 5 August 2015, Erio Camporesi, IT2574 (MFLU 15-2603, holotype), ex-type living culture (MFLUCC 18-0550). Culture characteristics: Colonies growing on PDA at 25 C reaching 2 cm in 1 week, irregular, white, fluffy appearance with undulate edge, reverse white. GenBank numbers: ITS: MK307843, LSU: MK307844, SSU: MK314097, TEF1-a: MK360181. Notes: In our phylogenetic analysis, Angustimassarina sylvatica groups as a distinct clade with A. alni, A. premilcurensis and Exosporium stylobatum (Fig. 9). Angustimassarina sylvatica has longer asci (95–110 lm) and longer ascospores (21–25 lm) than A. alni (asci 71–89 lm, ascospores 19–22 lm) and A. premilcurensis (asci 64–93 lm, ascospores 19–23 lm (Tibpromma et al. 2017). Additional details of morphological differences are mentioned in Table 1. Angustimassarina sylvatica was isolated from Fagus sylvatica (Sapindaceae) in Italy, whereas A. alni was recorded from Alnus glutinosa (Betulaceae) in Germany and A. premilcurensis recorded on a dead branch of Carpinus betulus (Betulaceae) in Italy (Tibpromma et al. 2017). There are six base pair differences between Angustimassarina sylvatica and A. alni and four base pair differences between A. sylvatica and A. premilcurensis for 485 ITS nucleotide sequences. There is no TEF1-a sequence data available for A. alni and A. premilcurensis. Camarosporidiellaceae Wanas., Wijayaw., Crous & K.D. Hyde Notes: In an investigation of camarosporium-like taxa, Wanasinghe et al. (2017a) introduced Camarosporidiellaceae to accommodate the genus Camarosporidiella which was well-positioned within the suborder Pleosporineae and phylogenetically distinct from other families. Camarosporidiellaceae species are characterized by coelomycetous asexual morphs, comprising pycnidial conidiomata, with a single, papillate ostiole, enteroblastic, annellidic, integrated to discrete, doliiform, lageniform or cylindrical, hyaline conidiogenous cells and pale to dark brown conidia that are phragmosporous to muriform and mostly ellipsoidal. Their sexual morphs are gregarious to solitary, globose to subglobose ascomata having a papillate, central ostiole, a peridium containing cell layers of textura angularis, cylindrical, (2–)4–8-spored asci and uniseriate, ellipsoidal, brown, muriform ascospores. However, taxonomic classification of Camarosporidiellaceae 123 18 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 7 Phylogram generated from maximum likelihood analysis based on combined LSU, SSU and TEF1-a sequence data. Seventysix strains are included in the combined analyses which comprised 2640 characters (804 characters for LSU, 995 characters for SSU, 841 characters for TEF1-a) after alignment. Hysterium angustatum (CBS 236.34) and (MFLU 16-1179) (Hysteriaceae, Hysteriales) are used as the outgroup taxa. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood value of - 19536.000988 is presented. Estimated base frequencies were as follows: A = 0.244757, C = 0.246173, G = 0.275709, T = 0.233360; substitution rates AC = 0.971666, AG = 3.423862, AT = 1.366124, CG = 1.101326, CT = 8.907769, GT = 1.000000; gamma distribution shape parameter a = 0.478984. Bootstrap support values for ML greater than 75% and Bayesian posterior probabilities greater than 0.95 are given near nodes respectively. The tree is rooted with H. angustatum (MFLU 16-1179 and CBS 236.34). Ex-type strains are in bold and black. The newly generated sequences are indicated in blue species based exclusively on morphological characteristics is insufficient for designating new species because of their inadequate phenotypic variability. Mostly taxa are saprobic and sometimes can be endophytic or pathogenic or potential opportunistic pathogens on leaves and woody materials. The current understanding of ecological and pathogenic aspects of this group is superficial and extensive sampling should be carried out in different regions and hosts. Pathogenetic virulence of this family also should be further investigated with more taxon sampling and DNA based sequence analyses. Camarosporidiella Wanas., Wijayaw., K.D. Hyde Notes: Camarosporidiella was established by Wanasinghe et al. (2017a) with C. caraganicola (Phukhams., Bulgakov & K.D. Hyde) Phukhams., Wanas. & K.D. Hyde as the type species. Wanasinghe et al. (2017a) placed 20 species in this genus based on morphological and multigene phylogenetic anlyses. In a recent study, Hyde et al. (2018b) added Camarosporidiella populina as a novel species from Russia. Species of Camarosporidiella are morphologically less-diverse and distributed throughout a broad range of environments especially in the Northern Hemisphere. Camarosporidiella species are saprobes, endophytes or pathogens of a wide range of hosts (Wanasinghe et al. 2017a). Species identification based solely on morphology and plant host association is difficult since many species have overlapping characters. The genus is similar to other camarosporium-like genera and distinguishing these genera can be problematic (Wanasinghe et al. 2017a). LSU is useful for preliminary genus identification. It is recommended to use a combination of LSU, SSU, ITS and TEF1-a sequence data (Wanasinghe et al. 2017a). In this study, we introduce the 22nd species of this 19 genus, Camarosporidiella mori, which was collected on Morus alba in Russia. Camarosporidiella mori Phutthacharoen, Bulgakov, Wanas. & K.D. Hyde, sp.nov. Index Fungorum number: IF555709; Facesoffungi number: FoF0549; Fig. 10 Etymology: The specific epithet ‘‘mori’’ refers to the plant host genus Morus. Holotype: MFLU 17–2147 Necrotrophic on branches of Morus alba L. Sexual morph Undetermined. Asexual morph Conidiomata 370– 520 9 220–430 lm ( x = 428 9 358 lm, n = 10), pycnidial, solitary or gregarious, scattered on wood, immersed to semi-erumpent, unilocular, dark brown, globose, black in centre with a papillate ostiole; Pycnidial wall 30–60 lm ( x = 38.5 lm, n = 10), composed of cells of textura angularis, multi-layered, with outer layer composed of 4–6 layers of heavily pigmented, light brown cells, inner layer composed of 3–4 layers, inwardly lighter. Conidiophores reduced to conidiogenous cells. Conidiogenous cells hyaline, rough, enteroblastic, numerous, clavate, formed from the inner most layer of the pycnidial wall. Conidia 15.5– 21 9 6.5–9.5 lm ( x = 18 9 8.5 lm, n = 20), hyaline when immature, brown when mature, oval, with 4-transverse septa and 1-longitudinal septum, smooth and thinwalled, blunt at both ends. Material examined: RUSSIA, Rostov region, Shakhty City, trees near buildings, on dying branch of Morus alba L. (Moraceae), 9 April 2017, Timur S. Bulgakov (MFLU 17–2147, holotype). GenBank numbers: ITS: MK590359, LSU: MK590358. Notes: Three species have been found on Morus alba, viz. Camarosporidiella celtidis, C. moricola and C. mori. Conidiomata are similar in dimensions and shape. They are unilocular, globose and black, with papillate ostioles in the centre. The shape of conidiogenous cells are not significantly differet. The conidia of C. mori are pale brown, while C. moricola has much darker conidia. Camarosporidiella mori conidia are comparatively larger than C. moricola (18 9 8.5 lm vs 11 9 5.5 lm). Camarosporidiella moricola and C. mori have conidia with longitudinal septa, while septa are lacking in C. celtidis. In the phylogenetic analyses, C. mori groups with C. eufemiaea and C. premilcurensis with high bootstrap support and is clearly distinct from C. moricola (Fig. 11). Comparison with the C. eufemiaea and C. premilcurensis asexual morph characters are currently impossible as they have introduced only as sexual morphs, while C. mori is known only from its asexual morph. Therefore, we introduce our novel collection as a new species. 123 20 Fungal Diversity (2019) 96:1–242 Fig. 8 Angustimassarina sylvatica (MFLU 15–2603, holotype). a Ascomata on the host material. b, c Sections through ascomata. d Peridium. e Pseudoparaphyses. f–h Ascospores. i–l Asci. Scale bars: b, c = 80 lm, d, i–l = 50 lm, f–h = 10 lm 123 Fungal Diversity (2019) 96:1–242 Cucurbitariaceae G. Winter Notes: The family Cucurbitariaceae, in the order Pleosporales, was established by Winter (1885) and Cucurbitaria Gray is the type genus (Doilom et al. 2013; Hyde et al. 2013; Wijayawardene et al. 2014; Li et al. 2016). Species belonging to this family have grouped, ostiolate ascomata located superficially on their hosts or on a pseudostromatic structure with cylindrical asci and phragmosporous or muriform ascospores (Hyde et al. 2013). They are either saprobes or necrotrophs on woody substrates (Hyde et al. 2013). While Index Fungorum (2019) lists numerous epithets belonging to Cucurbitariaceae, many of them are not linked to the latter and only a few species have molecular data in GenBank (Wanasinghe et al. 2017b; Jaklitsch et al. 2018). The family has been subjected to several generic additions and exclusions (Barr 1987; de Gruyter et al. 2010, 2013; Hyde et al. 2013; Doilom et al. 2013). Known asexual morphs of Cucurbitariaceae are pyrenochaeta- or phoma-like (Jaklitsch et al. 2018). For this reason, Pyrenochaeta De Not. was accommodated in the family (de Gruyter et al. 2010, 2013; Chen et al. 2015). Pyrenochaeta quercina Kabát & Bubák, P. unguis-hominis Punith. & M.P. English, P. cava (Schulzer) Gruyter, Aveskamp & Verkley, P. hakeae Crous and P. keratinophila Verkley, C. Ferrer & Gené (Wanasinghe et al. 2017b; Valenzuela-Lopez et al. 2018) were, however, recently transferred to the novel genus Neocucurbitaria Wanas., E.B.G. Jones & K.D. Hyde introduced by Wanasinghe et al. (2017b). Valenzuela-Lopez et al. (2018) introduced another genus, Paracucurbitaria Valenz.-Lopez, Stchigel, Guarro & Cano, which currently has two species, Paracucurbitaria corni (Bat. & A.F. Vital) Valenz.-Lopez, Stchigel, Guarro & Cano., formerly identified as Pyrenochaeta corni (Bat. & A.F. Vital) Boerema, Loer. & Hamers (Boerema et al. 1996) and a new species, P. italica Valenz.-Lopez, Crous, Stchigel, Guarro & Cano to the family. They established the new monotypic genus Allocucurbitaria Valenz.-Lopez, Stchigel, Guarro & Cano and the new families Pseudopyrenochaetaceae Valenz.Lopez, Crous, Stchigel, Guarro & Cano and Neopyrenochaetaceae Valenz.-Lopez, Crous, Cano, Guarro & Stchigel to accommodate taxa previously belonging to the genus Pyrenochaeta, thereby, eliminating Pyrenochaeta sensu stricto from Cucurbitariaceae. Additionally, species of Pyrenochaetopsis Gruyter, Aveskamp & Verkley, also previously included in Cucurbitariaceae, have been transferred to the new family Pyrenochaetopsidaceae Valenz.Lopez, Crous, Cano, Guarro & Stchigel (Valenzuela-Lopez et al. 2018). The latest generic additions to the family Cucurbitariaceae include Astragalicola Jaklitsch & Voglmayr, Cucitella Jaklitsch & Voglmayr, Parafenestella Jaklitsch & Voglmayr, Protofenestella Jaklitsch & 21 Voglmayr and Seltsamia Jaklitsch & Voglmayr (Jaklitsch et al. 2018). Protofenestella Jaklitsch & Voglmayr Notes: Protofenestella (Cucurbitariaceae, Pleosporales) was established and defined as the primitive form of the genus Fenestella Tul. & C. Tul. (Jaklitsch et al. 2018), by virtue of non-clustered, ill-defined ascomata on host materials, as well as distant phylogenetic placement. The genus currently accommodates the single species Protofenestella ulmi Jaklitsch & Voglmayr, whose sexual morph was reported on Ulmus sp. from Europe (Jaklitsch et al. 2018). Its asexual morph was described from culture (Jaklitsch et al. 2018). In this study, we introduce the asexual morph of P. ulmi from its host with evidence from morphology and concatenated phylogenetic analysis based on ITS, LSU and TEF1-a sequence data. Protofenestella ulmi Jaklitsch & Voglmayr Facesoffungi number: FoF05078; Fig. 12 Saprobic on dead twigs and branches of Ulmus pumila L. Sexual morph See Jaklitsch et al. (2018) for description and illustrations. Asexual morph Coelomycetous, phomalike. Conidiomata 78–131 lm high, 71–135 lm diam. ( x = 96 9 96.4 lm, n = 6), globose to depressed subglobose, erumpent through host periderm, solitary or loosely aggregated in ill-defined groups. Peridium 9–12.8 lm wide at the top, 5.8–8.9 lm wide at the sides, consisting of 4–6 layers, outer layer comprising olivaceous brown, relatively thin-walled cells of textura angularis, inner layer becoming paler and comprising pale brown, thin-walled cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 9–27 9 0.9–2.8 lm ( x = 14.3 9 1.66 lm, n = 32) hyaline, enteroblastic, phialidic, lageniform, doliiform to cylindrical, straight to faintly curved, discrete or integrated. Conidia 2.5– 5.4 9 0.9–2.4 lm ( x = 3.9 9 1.4 lm, n = 76), hyaline, aseptate, cylindrical or oblong to allantoid, mostly containing 2 small guttules, smooth-walled. Culture characteristics: Colonies on MEA reaching 1.5 cm diam. after 21 days at 25 C, circular, colony initially greyish olive, eventually turning unevenly brown after 4 weeks, with limited aerial mycelium, reverse olivaceous dull green to grey, odour indistinct. Material examined: RUSSIA, Rostov region, Shakhty City, Cotton fabric microdistrict, artificial grove (47.72195 N, 40.25177 E), on dead twigs of Ulmus pumila L. (Ulmaceae), 26 May 2017, Timur S. Bulgakov (MFLU 17-2047, HKAS 104976), living culture (MFLUCC 18-0376). GenBank numbers: ITS: MK418793, LSU: MK418783, TEF1-a: MK424969. Notes: The asexual morph of Protofenestella ulmi has so far only been reported from culture (Jaklitsch et al. 2018). 123 22 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 23 b Fig. 9 Phylogram generated from maximum likelihood analysis based on combined LSU, SSU, ITS and TEF1-a sequence data. Related sequences are obtained from GenBank. Melanomma pulvispyrius (CBS 124080) is used as the outgroup taxon. Twenty-nine strains are included in the combined analyses which comprise 3297 characters (525 characters for ITS, 847 characters for LSU, 984 characters for SSU, 941 characters for TEF1-a) after alignment. Single gene analyses were carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis. The best RaxML tree with a final likelihood value of - 12251.741088 is presented. The matrix had 784 distinct alignment patterns, with 18.70% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.244380, C = 0.247658, G = 0.268826, T = 0.239136; substitution rates AC = 1.50180, AG = 2.768806, AT = 1.335846, CG = 1.292688, CT = 6.996209, GT = 1.000000; gamma distribution shape parameter a = 0.137453. Maximum parsimony analysis of 2585 constant characters and 485 informative characters resulted in two equally most parsimonious tree of 1370 steps (CI = 0.688, RI = 0.838, RC = 0.577, HI = 0.312). Bootstrap support values for ML (first set) and MP equal to or greater than 50% are given above the nodes. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue In the present study, the asexual morph of the species is illustrated from natural substrate, Ulmus pumila. The strain obtained from this study clusters with the type strain of P. ulmi with high booststrap support (100% ML, 100 BYPP) in Fig. 13. The conidia of P. ulmi observed in our study are similar in size to those obtained from the ex-holotype culture [2.5–5.4 9 0.9–2.4 lm (this study) vs 2.9–5.4 9 1.3–2.2 lm (Jaklitsch et al. 2018)]; slight differences may be accounted for by the fact that the asexual morph of P. ulmi in this study was observed from natural substrate while that reported by Jaklitsch et al. (2018) was obtained from culture. Dictyosporiaceae Boonmee & K.D. Hyde Notes: The family Dictyosporiaceae was erected by Boonmee et al. (2016) with Dictyosporium Corda. as the type genus with morphological data and multi-gene analysis. The asexual morphs are cheirosporous hyphomycetes (Boonmee et al. 2016) and the sexual morphs are characterized by immersed to erumpent or superficial, globose to subglobose, dark brown to black ascomata, bitunicate asci and septate, hyaline, sheathed ascospores. Currently, there are twelve genera in this family (Wijayawardene et al. 2018a). Dictyocheirospora D’souza et al. Notes: Dictyocheirospora was introduced by Boonmee et al. (2016) with D. rotunda D’souza, Bhat & K.D. Hyde as the type species. Dictyocheirospora species are characterized by dark sporodochial colonies with aeroaquatic cheiroid dictyospores and all are saprobes (Boonmee et al. 2016). Currently, there are 19 epithets in Index Fungorum (2019). Dictyocheirospora aquadulcis Sorvongxay, S. Boonmee & K.D Hyde, sp. nov. Index Fungorum number: IF556308; Facesoffungi number: FoF05963; Fig. 14 Etymology: From the Latin aquadulcis (aqua = water and dulcis = sweet, in reference to its freshwater habitat. Holotype: MFLU 18-1088 Saprobic on decaying, submerged wood in freshwater. Sexual morph Undetermined. Asexual morph Hyphomycetous. Colonies on natural substrate, superficial, scattered. Mycelium immersed, branched, septate, pale brown, smooth-walled. Conidiomata 117–412 lm diam., slightly, pulvinate, dark brown. Conidiophores micronematous, undifferentiated from vegetative hyphae. Conidiogenous cells, holoblastic, integrated, terminal, pale brown. Conidia 60–80 lm 9 17–29 lm ( x = 70 9 24 lm, n = 20), solitary, cheiroid, ellipsoid to cylindrical, rounded at the apex, consisting of 7 rows of cells, with each row composed of 11–14 cells, with rows digitate, cylindrical, inwardly curved at apex, arising from Table 1 Synopsis of Angustimassarina species Species name Asci (lm) Ascospores (lm) Plant host Locality Reference A. acerina 92–105 21–23 Acer platanoides Germany Thambugala et al. (2015) A. alni 71–89 19–22 Alnus glutinosa Germany Tibpromma et al. (2017) A. arezzoensis 67–95 19–21 Salvia sp. Italy Tibpromma et al. (2017) A. coryli 70–100 20–25 Corylus avellane Italy Hyde et al. (2016) A. italica 78–103 15–22 Ilex aquifolium Italy Tibpromma et al. (2017) A. lonicerae 55–81 19–25 Lonicera sp. Italy Tibpromma et al. (2017) A. populi 80–95 19–22 Populus sp. Italy Thambugala et al. (2015) A. premilcurensis 64–93 19–23 Carpinus betulus Italy Tibpromma et al. (2017) A. quercicola 60–94 17–21 Quercus robur Germany Thambugala et al. (2015) A. rosarum 40–102 16–22 Rosa canina Italy Wanasinghe et al. (2018a) A. sylvatica 95–110 21–25 Fagus sylvatica Italy This study 123 24 Fungal Diversity (2019) 96:1–242 Fig. 10 Camarosporidiella mori (MFLU 17-2147, holotype). a Herbarium package. b Conidiomata on Morus alba. c Close up of conidiomata. d Cross section of conidioma. e Section through conidioma wall. f Conidiogenous cells. g–m Conidia. Scale bars: c = 1000 lm, d = 100 lm, e = 30 lm, f, g = 10 lm, h–m = 5 lm a basal cell, euseptate, slightly constricted at septa, brown, smooth-walled. Culture characteristics: Conidia germinated on WA within 24 h. Colonies on MEA reaching 0.2–0.5 cm diam., in 5 days at 25 C, with wavy margins, initial white and later becoming gray. Material examined: THAILAND, Pha Yao Province, on decaying wood, in flowing freshwater stream, 11 September 2017, Saranyaphat Boonmee (PK12), (MFLU 18-1088, holotype), ex-type living culture (MFLUCC 17-2571). GenBank numbers: ITS: MK634542, LSU: MK634545. Notes: Dictyocheirospora aquadulcis is similar to D. heptaspora in having the same conidial shape and size. However, it differs in having elongate conidia with 11–14 septa, whereas D. heptaspora has ellipsoidal conidia with 10–11 septa. Phylogenetically, D. aquadulcis clustered with D. heptaspora with moderate bootstrap support (Fig. 18). We therefore, introduce D. aquadulcis as a new species in this study based on its morphological distinctness. 123 Fungal Diversity (2019) 96:1–242 25 Fig. 11 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU and SSU sequence data for Camarosporidiella species and several closely related genera in Camarosporidiellaceae. Twenty-seven strains are included in the combined analyses which comprise 2376 characters (including gaps) (548 characters for ITS, 850 characters for LSU, 974 characters for SSU) after alignment. Coniothyrium palmarim (CBS 758.73 and CBS 400.71) (Coniothyriaceae, Pleosporales) are used as the outgroup taxa. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis was similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of - 4091.750594 is presented. The matrix had 130 distinct alignment patterns, with 5.97% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.252474, C = 0.213418, G = 0.274008, T = 0.260100; substitution rates AC = 3.818489, AG = 8.876577, AT = 3.746626, CG = 0.477392, CT = 7.229081, GT = 1.000000; gamma distribution shape parameter a = 0.020000. Maximum parsimony analysis of 2313 constant characters and 47 informative characters (CI = 0.607, RI = 0.756, RC = 0.459, HI = 0.393). Bootstrap values for maximum likelihood and maximum parsimony equal to or greater than 50 and Bayesian posterior probabilities equal or greater than 0.95 are placed above the branches respectively. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Dictyocheirospora rotunda M.J. D’souza, Bhat & K.D. Hyde Facesoffungi number: FoF05063; Fig. 15 Saprobic on dead wood. Sexual morph Undetermined. Asexual morph Hyphomycetous. Conidiomata punctiform, sporodochial, velvety, brown to dark brown. Mycelium superficial, with brown, smooth, septate, branched hyphae. Conidiophores micronematous, pale brown. Conidiogenous cells 5–8 9 6–9 lm ( x = 6.5 9 7 lm, n = 20), holoblastic, integrated, terminal, pale brown, smooth-walled. Conidia 16–29 9 47–62 lm ( x = 22 9 55 lm, n = 50), solitary, monoblastic, acrogenous, cheiroid, pale brown to brown, consisting of 6–8 rows of cells, rows digitate, cylindrical, inwardly curved at the tip, arising from a basal cell, each arm composed of 7–9-septated, constricted at septa. Culture characteristics: Ascospores germinating on PDA (potato dextrose agar) within 1 week at 23 C. Sporulation after 30 days, colony 1.5 cm diam. Initially, orange from above and reverse, becoming brown, with concentric zonation, with irregular form of margin, filamentous mycelium, with rough surface and raised elevation. Material examined: CHINA, Yunnan Province, Kunming, Songhuaba Lake, on dead wood, 3 September 2017, 123 26 Fig. 12 Protofenestella ulmi (MFLU 17-2047, new record). a, b Appearance of conidiomata on host substrate. c Vertical section through conidioma. d Peridium. e Conidiogenous cells. f Mature and immature conidia attached to conidiogenous cells. g–i Conidia. j, k 123 Fungal Diversity (2019) 96:1–242 Colony on MEA (j from above view, k from below view). l Conidiomata on colony. m Conidia attached to conidiogenous cell in culture. Scale bars: a = 1000 lm, b, l = 100 lm, c = 50 lm, d, e = 20 lm, f = 10 lm, m = 7 lm, g–i = 3 lm Fungal Diversity (2019) 96:1–242 S.K. Huang (KUN-HKAS 99571), living culture (KUMCC 18-0014). GenBank numbers: ITS: MK050008, LSU: MK050010, SSU: MK050009. Notes: Dictyocheirospora rotunda has been previously reported from Thailand and northwestern Yunnan Province, China (Boonmee et al. 2016; Wang et al. 2016). In phylogenetic analysis of combined ITS and LSU sequence data of Dictyosporiaceae strains, our strain clusters with the ex-type strain of Dictyocheirospora rotunda (MFLUCC 14-0293) with relatively high-support (Fig. 18). Pairwise comparison of DNA sequences of ITS and LSU regions between KUMCC 18-0014 and MFLUCC 14-0293 reveals they are slightly different. Boonmee et al. (2016) introduced D. rotunda as having 5–7 9 6–7 lm, pale brown conidiogenous cells and 42–58 9 19–38 lm, cylindrical conidia. The morphological similarities reveal they are the same species. Thus, a record of D. rotunda from Kunming, Yunnan Province is reported. Dictyocheirospora taiwanense Tennakoon, C.H. Kuo & K.D. Hyde, sp. nov. Index Fungorum number: IF556309; Facesoffungi number: FoF05964; Fig. 16 Etymology: Named after Taiwan, where this fungus was collected. Holotype: MFLU 18-0070 Saprobic on decaying wood of Macaranga tanarius (L.) Müll.Arg. Sexual morph Undetermined. Asexual morph Hyphomycetous. Conidiomata punctiform, sporodochial, 110–230 lm diam., velvety, dark brown to black. Conidiophores micronematous, pale brown, smooth, thin-walled. Conidiogenous cells holoblastic, cylindrical. Conidia (72–) 74–84(–86) 9 16–20(–24) lm ( x = 77.5 9 18 lm, n = 30), solitary, brown, ellipsoid to cylindrical, rounded at the apex, cheiroid, not complanate, consisting of 5 rows of cells, rows cylindrical, arising from a basal cell, each composed of 10–13 cells, constricted at septa, without appendages. Culture characteristics: Conidia germinating on water agar within 24 h, germ tubes appearing from the basal cells of the conidium. Colonies on PDA covering 3 cm diam., in 3 weeks at 27 C, at first white, colony from above: light orange to yellowish at the margin, yellowish to white at the centre; reverse, orange to yellowish at the margin, orange at the centre; mycelium yellowish to white with tufting; not producing pigments in PDA. Material examined: TAIWAN, Chiayi, Shihnong Forest Area, decaying wood of Macaranga tanarius (L.) Müll.Arg. (Euphobiaceae), 25 June 2017, D.S. Tennakoon, DTW013 (MFLU 18-0070, holotype), ex-type living culture (MFLUCC 17-2654). GenBank numbers: ITS: MK495821, LSU: MK495820. 27 Notes: Dictyocheirospora taiwanense is typical of Dictyocheirospora in having aeroaquatic cheiroid dictyospores (Boonmee et al. 2016). Phylogenetically, D. taiwanense nests independently closely to D. garethjonesii (MFLUCC 16-0909), D. aquatica (KUMCC 15-0305), D. cheirospora (KUMCC 17-0035) and D. rotunda (MFLUCC 14-0293a, HKAS 99571) with moderate support (Fig. 18). However, D. taiwanense differs from D. aquatica and D. garethjonesii in having larger conidia (74–84 9 16–20 lm) with a higher number of cells (10–13) in each conidial row, whereas the conidia of D. aquatica (34–42 9 12.5–19.5 lm) and D. garethjonesii (45.5–54.5 9 15.5–24.5 lm) are comparatively smaller and comprise 6–8, 7–10 cells in each conidial row (Wang et al. 2016). In addition, D. pseudomusae and D. vinaya are also show distinct morphological differences viz. mostly D. taiwanense conidia has 5 rows (10-13 cells in each row), whereas D. pseudomusae has 6-7 rows (13–15 cells in each row) and D. vinaya (6–7 rows, 9–13 cells in each row) (Tanaka et al. 2015; Boonmee et al. 2016). Digitodesmium Corda, Weitenweber’s Beitr. Notes: Digitodesmium was established by Kirk (1981) with type species Digitodesmium elegans P.M. Kirk. The asexual morph of the Digitodesmium species is characterized by punctiform, sporodochial conidiomata and acrogenous, euseptate, cheiroid, digitate conidia, with apical gelatinous caps. There are six records in Index Fungorum (2019), however, these are lacking the molecular data with the exception of D. bambusicola. Digitodesmium chiangmaiense Q.J. Shang & K.D. Hyde sp. nov. Indexfungroum number: IF556220; Facesoffungi number: FoF05098; Fig. 17 Etymology: Names reflects the province in Thailand, from where the holotype was collected. Holotype KUN-HKAS 102163 Saprobic on decaying wood submerged in stream. Sexual morph Undetermined. Asexual morph Hyphomycetous. Colonies punctiform, sporodochial, scattered, dark brown to black, glistening. Conidiophores micronematous, hyaline to pale brown, unbranched, thinwalled, smooth. Conidiogenous cells monoblastic, integrated, terminal, determinate, hyaline to pale brown, smooth. Conidia acrogenous, solitary, cheiroid, smoothwalled, complanate, brown to dark brown, consisting of three arms closely compact with side arms higher than middle arms, (24–)33–42(–44) 9 (13–)15–18(–21) lm ( x = 38 9 17 lm, n = 45), 5–7-euseptate in each arm, without appendages attached at the two outer arms, arms (20–)27–37(–38) 9 (4–)4–7(–8) lm ( x = 32 9 5 lm, n = 30), not discrete, unbranched, cylindrical. 123 28 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 13 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU and TEF1-a sequence data for Protofenestella species and several closely related genera in Cucurbitariaceae. Related sequences are taken from Wanasinghe et al. (2017b) and Jaklitsch et al. (2018). Forty-three strains are included in the combined analyses which comprise 2749 characters (601 characters for ITS, 914 characters for LSU and 1234 characters for TEF1a) after alignment. Pyrenochaetopsis leptospora (CBS 101635) (Pyrenochaetopsidaceae, Pleosporales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree has clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood value of - 13415.377769 is presented. The matrix had 741 distinct alignment patterns, with 25.51% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.234883, C = 0.253189, G = 0.270793, T = 0.241136; substitution rates AC = 1.125763, AG = 3.623108, AT = 2.100279, CG = 1.168787, CT = 6.351079, GT = 1.000000; gamma distribution shape parameter a = 0.142500. Bootstrap values for maximum likelihood equal to or greater than 60 are placed above and below the branches respectively. Branches with Bayesian posterior probabilities equal or greater than 0.90 are in bold. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Culture characteristics: Conidia germinating on PDA within 24 h. Germ tubes produced from the basal cell. Colonies on PDA reaching 3–5 mm diam. after 7 days at room temperature, colonies circular, with fluffy, dense, white mycelium in the middle and sparse mycelium in the outer ring on the surface with irregular brown margin; in reverse, brown in the middle and pale yellow to brown at the margin. Material examined: THAILAND, Chiang Mai Province, Mae Taeng District, Mushroom Research Center, on dead wood, 27 January 2017, Qiuju Shang, M7P4-17 (KUNHKAS 102163, holotype), ex-type living culture (MFLUCC, ICMP). GenBank numbers: LSU: MK571766, SSU: MK571775. Notes: Based on analysis of combined ITS and LSU sequence data (Fig. 18), Digitodesmium chiangmaiense (M7P4-17) groups with D. bambusicola (CBS 110279) with moderate support (57% ML, 52% MP and 0.92 BYPP). Digitodesmium chiangmaiense differs from D. bambusicola in lacking appendages on the outer of the arms and having longer conidia (Table 2). In addition, D. chiangmaiense differs from other Digitodesmium species in the dimension and number of arms (Table 2). Didymellaceae Gruyter, Aveskamp & Verkley Notes: The family Didymellaceae was established to accommodate Didymella, Phoma and phoma-like genera and is one of the most species rich families in the fungal kingdom (Chen et al. 2015, 2017), with 31 genera based on morphology and phylogeny (Chen et al. 2015; Jayasiri et al. 2017; Thambugala et al. 2017a; Wanasinghe et al. 2018b). Species of this family are distributed in a broad 29 range of environments. Most members are economically important plant pathogens causing leaf and stem spots and many are mentioned in quarantine regulations (Thambugala et al. 2017b, 2018; Knight et al. 2018; ValenzuelaLopez et al. 2018). Limited studies have been carried out on the sexual morphs (Chen et al. 2015; Thambugala et al. 2017a, 2018; Wanasinghe et al. 2018b). The taxonomic placements of this family were treated based on phylogenetic and morphological analyses (Chen et al. 2015; Jayasiri et al. 2017; Thambugala et al. 2017a; Wanasinghe et al. 2018b). Calophoma Qian Chen & L. Cai Notes: The genus Calophoma, introduced by Chen et al. (2015), and is typified by C. clematidina. (Thum.) Q. Chen & L. Cai and accommodates C. aquilegiicola, C. clematidina, C. clematidis-rectae, C. complanata, C. glaucii and an insufficiently known species (CBS 186.55) (Chen et al. 2015). The sexual morph has been reported for C. petasitis (Tibpromma et al. 2017). The asexual morph has been reported with its unique characteristics such as subglobose, subcylindrical, ellipsoidal, somewhat obclavate-fusiform conidia with 0–1 septa and chlamydospores are produced in one species (Chen et al. 2015, 2017; Tibpromma et al. 2017). Currently there are eight species included in this genus (Tibpromma et al. 2017). We introduce a novel species to this genus with support from combined LSU, ITS, TUB2 and RPB2 sequence data. Calophoma humuli Thiyagaraja., Bulgakov & K.D. Hyde, sp. nov. Index Fungorum number: IF555424; Facesoffungi number: FoF05077; Fig. 19 Etymology: The specific epithet ‘‘humuli’’ refers to the host plant genus Humulus. Holotype: MFLU 17-2205 Saprobic on dead stems of Humulus lupulus L. (Cannabaceae). Sexual morph Undetermined. Asexual morph Coelomycetous. Conidiomata 190–220 lm high 9 210– 230 lm diam. ( x = 205 9 220 lm, n = 5), pycnidial, subimmersed, solitary, globose, black. Pycnidial wall multilayered, 10–25 lm wide at the base, 10–20 lm wide in sides, thick, comprising two layers, outer layer heavily pigmented, thick walled, comprising brown cells of texura angularis, cells towards the inside lighter, inner layer composed of hyaline texura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells phialidic, hyaline, smooth walled, formed from inner layer of the pycnidium wall. Conidia 3–10 9 3–6 lm ( x = 6.5 9 3.8 lm, n = 50), oblong, straight, rounded at the ends, sometimes narrowly rounded at the ends, smooth and thick-walled, aseptate when immature and becoming 1-septate when mature, initially hyaline, becoming pale brown at maturity. 123 30 Fungal Diversity (2019) 96:1–242 Fig. 14 Dictyocheirospora aquadulcis (MFLU 18–1088, holotype). a, b Colonies on host surface. c Colony of conidia. d–m Conidium arms. n Germinating conidium. o, p Culture on MEA at 3 weeks. q, r Culture on MEA at 7 months. Scale bars: a, b = 200 lm, c = 100 lm, e–m = 50 lm, d, n = 20 lm Culture characteristics: Colonies on PDA reaching 40–45 mm diam. after 2 weeks at 20–25 C, dense, circular, slightly raised, surface smooth with even edge, cottony, colony from above: white at the margin, grayish at the centre; from below: white at the margin, brownish yellow at the centre; not producing pigmentation on PDA media. Material examined: RUSSIA, Rostov region, Krasnosulinsky District, Gornensky protected landscape, edge of ravine forest (47.86602 N, 40.22959 E), on dead stems of 123 Fungal Diversity (2019) 96:1–242 31 Fig. 15 Dictyocheirospora rotunda (HKAS 99571, new geographical record). a Material. b Appearance of conidiomata on host. c Conidioma. d–f Conidia. g Squashed conidium. h–j Arms of conidia. Scale bars: d–g = 20 lm, h–j = 10 lm Humulus lupulus L. (Cannabaceae), 6 July 2017, Timur S. Bulgakov (MFLU 17-2205, holotype), ex-type living culture (MFLUCC 18-0101). GenBank numbers: ITS: MK446317, LSU: MK446318, RPB2: MK492667, TUB2: MK492666. Notes: In our multi-gene analyses, Calophoma humuli forms a monophyletic clade with C. petasitis with strong support. Calophoma petasitis is reported as the sexual morph in the genus Calophoma with the characteristics of subglobose ascomata and smooth-walled, guttulate, hyaline ascospores. A comparison of LSU, ITS and RPB2 with these two strains reveals 3/820 (0.36%), 9/441 (2.05%) and 75/797 (9.4%) base pair differences. Even though their LSU and ITS sequences are quite similar, the RPB2 region reveals they should be two different species following the guidelines of Jeewon and Hyde (2016). Therefore, we introduce our isolate as a new species in Calophoma. Neodidymelliopsis Qian Chen & L. Cai Notes: Neodidymelliopsis was introduced by Chen et al. (2015), with Neodidymelliopsis cannabis (G. Winter) Q. Chen & L. Cai as the type species (Chen et al. 2015; Hyde et al. 2016). 123 32 Fungal Diversity (2019) 96:1–242 Fig. 16 Dictyocheirospora taiwanense (MFLU 18–0070, holotype). a Conidiomata on the substrate. b, c Close-up of conidiomata. d Squash mount of conidia. e, f Squash mount of conidioma with conidiogenous cells. g–j Conidia. k Germinating conidium. Scale bars: d = 100 lm, e = 50 lm, f–k = 20 lm Neodidymelliopsis farokhinejadii S.A. Ahmadpour & M. Mehrabi-Koushki, Sydowia 69: 175 (2017) Facesoffungi number: FoF04975; Fig. 20 Saprobic or necrotrophic on dead arial stem of Cirsium sp. Sexual morph Undetermined. Asexual morph Coelomycetous. Conidiomata on host 75–105 lm diam. ( x = 74 lm, n = 10), pycnidial, solitary, scattered, globose to subglobose, black, semi-immersed to immersed; pycnidial wall pseudoparenchymatous, textura angularis dark brown to hyline cells. Chlamydospores not observed. Conidiophores not observed. Conidiogenous cells not observed. Conidia on host 4.5–7.2 9 2.3–3.6 lm diam. ( x = 6 9 3.2 lm, n = 40), oblong to ellipsoid, hyaline, aseptate, smooth-walled; in culture, 4.6–7.5 9 2.4–3.9 lm diam. ( x = 6 9 3.4 lm, n = 40), ellipsoidal to cylindrical, sometimes allantoid, hyaline, smooth- and thin-walled, aseptate, with small polar guttules. Culture characteristics: Colonies on PDA reach 25–30 mm diam. after 7 days at 25 C, with white aerial mycelium, surface floccose to wooly, margin regular, pale greenish olivaceous; reverse dull green. Material examined: ITALY, Forlı̀-Cesena Province, near Castrocaro Terme, on dead aerial stem of Cirsium sp. (Asteraceae), 25 February 2018, Erio Camporesi, living culture (MFLUCC 18-1569). Genbank numbers: ITS: MK084580, LSU: MK084581. Notes: Neodidymelliopsis farokhinejadii was first reported on dead branches of Eucalyptus sp. and was subsequently found on Citrus limon, Conocarpus erectus, Ziziphus sp., Juglans regia and Cupressus sp. (Ahmadpour et al. 2017) Our new isolate of N. farokhinejadii was collected from Forlı̀-Cesena Province in Italy. This is the first report of N. farokhinejadii on Cirsium species (Farr and Rossman 2019). Morphological observations such as spore characteristics support the inclusion of our taxa into Neodidymelliopsis (Chen et al. 2015). DNA sequence analyses from the LSU, ITS, TUB2 and RPB2 genes confirms that our taxon clades together with N. farokhinejadii (Fig. 21). 123 Fungal Diversity (2019) 96:1–242 Fig. 17 Digitodesmium chiangmaiense (KUN-HKAS 102163, holotype). a, b Colonies on submerged wood. c Squash mount of a sporodochium. d–g Conidia with conidiophores. h Conidium. 33 i Germinated conidia. j, k Culture characteristics on PDA (j from above view, k from below view) Scale bars: c = 20 lm, d–i = 10 lm 123 34 Fungal Diversity (2019) 96:1–242 Fig. 18 Phylogram generated from maximum likelihood analysis based on combined ITS and LSU sequences data in Dictyosporiaceae. Related sequences are taken from Boonmee et al. (2016) and Wang et al. (2016). Fifty-three strains are included in the combined analyses which comprise 1405 characters (553 characters for ITS, 852 characters for LSU) after alignment. Letendraea helminthicola (7-3) (Montagnulaceae, Pleosporales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis and the maximum parsimony analysis. The best RAxML tree with a final likelihood value of - 8736.227244 is presented. The matrix had 523 distinct alignment patterns, with 18.97% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.241508, C = 0.237486, G = 0.270256, T = 0.250751; substitution rates AC = 2.173342, AG = 3.473225, AT = 3.246966, CG = 0.531302, CT = 36.318285, GT = 1.000000; gamma distribution shape parameter a = 0.567927. Maximum parsimony analysis of 981 constant characters and 292 informative characters resulted in two equally most parsimonious trees of 1392 steps (CI = 0.486, RI = 0.744 RC = 0.361, HI = 0.514). Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 60 are placed above and below the branches respectively. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are in bold. The ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue Neodidymelliopsis sambuci Manawasinghe, Camporesi & K.D. Hyde, sp. nov. Index Fungorum number: IF555378; Facesoffungi number: FoF04939; Fig. 22 Etymology: Name reflects the host genus. Holotype: MFLU 18-0177 Saprobic or necrotrophic on dead and dying twigs and branches of Sambucus nigra. Sexual morph Undetermined. Asexual morph Coelomycetous. Conidiomata on host 37–82 lm diam. ( x = 58 lm, n = 10), pycnidial, solitary, scattered, globose, black, semi-immersed to immersed. Conidiophores not observed. Conidiogenous cells not observed. Conidia on host 4–7 9 2–3.5 lm diam. ( x = 6 9 2 lm, n = 40), oblong to ellipsoid, hyaline, aseptate, smooth-walled; in culture, 5–8 9 1–3.5 lm diam. ( x = 6 9 2.7 lm, n = 40). Conidial exudates not observed.Culture characteristics: Colonies on PDA reach 65 mm diam. after 7 days at 25 C, with circular, entire edge, raised grey olivaceous aerial mycelium, surface floccose to woolly. Material examined: ITALY, Province of Ravenna, Santa Lucia, Faenza, on living branch of Sambucus nigra L. (Adoxaceae), 5 January 2018, Erio Camporesi (MFLU 123 Fungal Diversity (2019) 96:1–242 35 Table 2 Conidia of Digitodesmium species discussed in this study Taxa D. bambusicola D. chiangmaiense D. elegans D. heptasporum D. intermedium D. macrosporum D. recurvum Colour Pale brown Brown to dark brown – Pale brown Brown to dark brown Brown to dark brown Pale brown Dimension (lm) 24– 32.5 9 12.5– 23 (25–)30–45(– 44) 9 (13– )12–21(–21) 45– 60 9 12– 21 50– 75 9 32.5– 70 39–76 9 25– 35 130– 145 9 19–26 30– 45 9 12.5– 23 Appendages Yes No No No No No No Wall Smooth Smooth Smooth Smooth Smooth Smooth Smooth Septal pores Conspicuous Conspicuous Inconspicous Conspicuous Conspicuous Conspicuous Conspicuous Number of arms 3 3 (2–)3–4(–6) 6–7 3–11 5–8 (2–)4–6(–7) Number of septa in each of arm 4–7 5–7 9–12 11–17 7–13 17–19 6–10 Country Philippines Thailand UK China Spain Spain China Reference Cai et al. (2002) This study Kirk. (1981) Cai et al. (2003) Silvera-Simón et al. (2010) Silvera-Simón et al. (2010) Ho et al. (1999) Fig. 19 Calophoma humuli (MFLU 17–2205, holotype). a, b Conidiomata on host. c Section through conidioma. d Section through conidioma wall. e Conidiogenous cells. f–i Conidia. Scale bars: a, b = 500 lm, c = 100 lm, d = 20 lm, e = 10 lm, f–i = 5 lm 123 36 Fungal Diversity (2019) 96:1–242 Fig. 20 Neodidymelliopsis farokhinejadii (JZB–H380023, new host record). a, b Conidiomata on the host tissue. c Section of the pycnidial wall. d–h Conidia. i Upper view of 7-day old culture. j Reverse view of 7-day old culture. Scale bar: c–h = 10 lm 18-0177, holotype), ex-type living culture (MFLUCC 18-1565). GenBank numbers: ITS: MH84232, LSU: MH84231, TUB2: MK049556. Notes: Morphological characters such as conidial size and shape (Fig. 22) fit well within the concept of Neodidymelliopsis (Chen et al. 2015). The combined gene sequence analyses of LSU, ITS, RPB2 and TUB2 genes confirm that present taxon belongs in genus Neodidymelliopsis in a strongly supported monophyletic lineage (Fig. 21) within family Didymellaceae. In phylogenetic analysis Neodidymelliopsis sambuci is in a sister clade with N. longicolla. Morphologically, N. sambuci can be distinguished from N. longicolla by its smaller conidia (12–15 9 4–7 lm in N. longicolla). This is the first report 123 of Neodidymelliopsis reported on Sambucus species (Farr and Rossman 2019). Nothophoma Qian Chen & L. Cai Notes: The genus Nothophoma was introduced by Chen et al. (2015) with the type species Nothophoma infossa (Ellis & Everh.) Q. Chen & L. Cai. (syn. Phoma infossa) and five other species transferred from Phoma, namely N. anigozanthi (syn. P. anigozanthi), N. arachidis-hypogaeae (syn. P. arachidis-hypogaeae), N. quercina (syn. P. fungicola), and N. gossypiicola (syn. P. gossypiicola) (Chen et al. 2015). There are nine species in this genus with the recently introduced N. variabilis (Valenzuela-Lopez et al. 2018). Ovoid, oblong to ellipsoidal aseptate conidia are a characteristic of the genus Nothophoma. This genus Fungal Diversity (2019) 96:1–242 37 Fig. 21 Phylogram generated from maximum likelihood analysis based on combined LSU, ITS, TUB2 and RPB2 sequence data for Didymellaceae. Related sequences are taken from Chen et al. (2015), Jayasiri et al. (2017), Thambugala et al. (2017a) and Wanasinghe et al. (2018b). Fifty-nine strains are included in the combined genes sequence analyses which comprise total 2683 characters (960 characters for LSU, 547 characters for ITS, 360 characters for TUB2, 816 characters for RPB2) after alignment. Leptosphaeria conidia (CBS 616.75) and L. doliolum (CBS 505.75) (Leptosphaeriaceae, Pleosporales) are used as the outgroup taxa. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best sorting RaxML tree with a final likelihood value of - 16885.291744 is presented. The matrix had 945 distinct alignment patterns, with 21.62% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.241571, C = 0.241856, G = 0.271818, T = 0.244754; substitution rates AC = 1.630628, AG = 4.058235, AT = 1.775626, CG = 0.954941, CT = 9.716323, GT = 1.000000; gamma distribution shape parameter a = 0.414994. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are in bold. The ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue 123 38 Fungal Diversity (2019) 96:1–242 Fig. 22 Neodidymelliopsis sambuci (MFLU 18-0177, holotype). a Appearance of pycnidia on dead branch of Sambucus nigari. b Appearance of pycnidia under the light microscope. submerged conidiomata on the host surface. c–e Conidia on host. f Pycnidia on agar. g pycnidial wall. h Developing hyphae. i–k Conidia on PDA. l Upper view of 7 days old culture on PDA. m Reverse view of 7 days old culture on PDA. Scale bars: a = 1000 lm, b, f = 100 lm, c, k = 20 lm, d, e, i, j = 10 lm contains ubiquitous fungal species, of which many are plant pathogens, some of which are of quarantine concern (Chen et al. 2015). Pathogenic on leaves of Buxus megistophylla H.Lév. Sexual morph Undetermined. Asexual morph Conidiomata produced on the PDA surface, 0.19–0.47 mm ( x = 0.25 mm, n = 10) diam., pycnidial, solitary, scattered, globose to irregularly shaped, black, ostiolate. Pycnidial wall multi-layered, composed of pale brown, Nothophoma quercina (Syd. & P. Syd.) Q. Chen & L. Cai, Stud. Mycol. 82: 213. 2015. Facesoffungi number: FoF04974; Fig. 23 123 Fungal Diversity (2019) 96:1–242 39 Fig. 23 Nothophoma quercina (JZB380024, new host record). a, b Conidiomata on the host tissue. c Section of the pycnidial wall. d Pycnidia on PDA. e Conidiogenous cells on culture. f Conidia on culture. g Upper view on colonies on PDA. h Lower view on colonies on PDA. Scale bar: c = 1 mm, d–f = 20 lm pseudoparenchymatous cells. Conidiogenous cells phialidic, hyaline, simple, doliiform to ampulliform, variable in size. Conidia 4–5.6 9 1.4–3.3 lm ( x = 5.2 9 2.3 lm, n = 40), variable in size and shape, subglobose to oval or obtuse, initially hyaline, but brown at maturity, aseptate, smooth-walled. Conidial exudates not recorded. Culture characteristics: Colonies on PDA reach 80 mm diam. after 7 days at 28 C, with regular margin, dull white aerial mycelium surface floccose to wooly, greenish olivaceous to olivaceous near the centre and reverse dark ochreous in the centre and white in the margin. Material examined: CHINA, Haidian District, Beijing, on leaves of Buxus megistophylla (Buxaceae), November 2017, D. Harishchandra, living culture (JZB380024, MFLUCC 18-1568). GenBank numbers: ITS: MK070136, LSU: MK070139. Notes: Nothophoma quercina has been reported as a pathogen on Pistacia vera (Chen et al. 2013), Chaenomeles sinensis (Yun and Oh 2016), Olea europaea (Moral et al. 123 40 2017) and was reported as a saprobe on Quercus sp. in Ukraine (Chen et al. 2015). This is the first record of N. quercina on Buxus sp. Phylogenetic analysis with the combined sequence data of ITS, TEF1-a, TUB2 and RPB2 of our strain (MFLUCC 18-1568) clusters together with the ex-type strain of Nothophoma quercina (CBS 633.92) with relatively high bootstrap and Bayesian probabilities (99% MP/1.00 PP). Hermatomycetaceae Locq. Notes: Hashimoto et al. (2017) introduced Hermatomycetaceae for a separate lineage of Hermatomyces species in Pleosporales, with Hermatomyces Speg. as the generic type. Hermatomycetaceae was introuduced by Locquin (1984) based on their distinctive characteristics, recently, the family placement had been formalised with a robust phylogenetic information (Doilom et al. 2017; Hashimoto et al. 2017). Several taxa have been introduced in the family, interestingly, no sexual morph has been discovered (Koukol et al. 2018; Tibpromma et al. 2016, 2018). In this study, we introduce a novel Hermatomyces species based on the distinctive morphology and supportive phylogenetic relationships (Fig. 24). Hermatomyces Speg. Notes: Hermatomyces was introduced by Spegazzini (1911) with H. tucumanensis Speg. as a type species. The genus is commonly present as sporodochial conidiomata and muriform, lenticular, hyaline or dematiaceous conidia of one or two types (Chang 1995; Tibpromma et al. 2016; Hashimoto et al. 2017). Koukol et al. (2018) synonymized several species of Hermatomyces based on morphological similarity and possible artefacts in the phylogenetic analysis. Tibpromma et al. (2018) confirm that the species homogeneity needs sufficient evidence from characters and TUB2 gene sequence data are still required (Jeewon and Hyde 2016). Therefore in this study, we follow Tibpromma et al. (2018) and introducing a novel species of Hermatomyces, H. bauhiniae based on its morphological and phylogenetically distinct from other Hermatomyces species (Fig. 24). Hermatomyces bauhiniae Phukhams., D.J. Bhat & K.D. Hyde, sp. nov. Index Fungorum number: IF555500; Facesoffungi number: FoF04827; Fig. 25 Etymology: Name refers to the host plant, Bauhinia variegata from which this fungus was isolated. Holotype: MFLU 18-1381 Saprobic on dead branch of Bauhinia variegata L. Sexual morph Undetermined. Asexual morph Colonies on natural substrate forming sporodochial conidiomata, superficial, scattered, circular or oval, blackish brown, velvety, glistening, consisting of an orbicular, abundantly 123 Fungal Diversity (2019) 96:1–242 sporulation, conidia readily liberated when agitated, 160– 180 lm wide. Mycelium mostly superficial, composed of a loose or compact network of repent, branched, septate, rough-walled, thick-walled, reddish brown to brown hyphae, 2–4 lm wide; subicular hyphae short, irregularly geniculate or flexuous, densely packed. Conidiophores 4– 10 9 2–5 lm, micronematous or semimacronematous, mononematous, cylindrical, erect, smooth or finely verruculose, aseptate, unbranched, often corresponding to conidiogenous cells, hyaline. Conidiogenous cells 3– 8 9 4–9 lm, holoblastic, monoblastic, integrated, terminal, determinate, cylindrical or slightly subulate, subsphaerical or ampulliform, hyaline. Conidia dimorphic, solitary, smooth-walled: lenticular conidia: 25–36 9 15– 20 lm ( x = 30 9 18 lm, n = 70), muriform, smooth, broadly ellipsoidal to oval in front view, brown to dark brown, slightly constricted at the septa, in side view composed of one column of 4–6 cells, end cells subhyaline to pale brown, often carrying remnant of conidiogenous cell at base; cylindrical conidia: 20–28 9 8–11 lm ( x = 24 9 9 lm, n = 30), straight or flexuous, septate, constricted at the septa, consisting of one column, 2–3septate, doliiform, cylindrical or subcylindrical, apical cells clavate or doliiform, verrucose, apex rounded, basal cells globose or subglobose, smooth, hyaline. Culture characteristics: Colonies on MEA at room temperature (25 8C) reaching 7 cm in 2 weeks, circular with entire margin, with white–grey mycelium, white at the margin, smooth at the surface and raised, mycelium strongly radiating into agar; reverse beige, no sporulation in cultures. Material examined: THAILAND, Phrae Province, on dried branch of Bauhinia variegata (Fabaceae), 25 July 2015, C. Phukhamsakda, S1_02 (MFLU 18-1381, holotype), ex-type living culture (MFLUCC 16-0395, ICMP 21948). GenBank numbers: ITS: MK443382, LSU: MK443378, RPB2: MK443386, SSU: MK443380, TEF1-a: MK443384. Notes: Hermatomyces bauhiniae is introduced as a new species based on its distinct morphology and support from phylogenetic analysis. Hermatomyces bauhiniae has distinct characteristics as compared to the other species. The lenticular conidia are not round or flat in one plane, and have well visible conidiogenous cells attached at the bottom of the conidia. In the phylogenetic analysis, H. bauhiniae formed a clade with H. indicus with good support (MLBS = 82; BYPP = 0.99). Both species were recorded from angiosperms plants. Nonetheless, H. bauhiniae can be distinguished from H. indicus by having straight conidia with only one column, hyaline, and cylindrical conidia, where as in H. indicus conidia are turbinate, grey-brown at the apex, and become hyaline towards the base with two Fungal Diversity (2019) 96:1–242 41 Fig. 24 Phylogram of 50% majority rule consensus tree from the Bayesian-inference analysis based on combined dataset of LSU, ITS, TEF1-a, and RPB2 sequence data representing Hermatomycetaceae. Related sequences are taken from Doilom et al. (2017), Hashimoto et al. (2017), Koukol et al. (2018) and Tibpromma et al. (2018). Thirty-nine strains are included in the combined analyses which comprise 3301 characters (826 characters for LSU, 514 characters for ITS, 941 characters for TEF1-a, 1020 characters for RPB2) after alignment. The tree is rooted with Anteaglonium globosum (ANM 925.2) and A. parvulum (MFLUCC 14-0821) in Anteagloniaceae (Pleosporales). Single gene analyses were also performed to compare the topology and clade stability with combined gene analyses. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. Bootstrap support values for maximum likelihood analysis greater than 70% and clade credibility values greater than 0.90 (the rounding of values to 2 decimal proportions) from Bayesianinference analysis labeled on the nodes. The matrix has 710 distinct alignment patterns, with 20.65% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.244863, C = 0.263683, G = 0.262259, T = 0.229195; substitution rates AC = 1.191383, AG = 5.252837, AT = 1.421242, CG = 1.059773, CT = 14.708510, GT = 1.000000; gamma distribution shape parameter a = 0.766883. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue columns. Hermatomyces indicus has larger cylindrical conidia but smaller lenticular conidia (Prasher and Sushma 2014; Doilom et al. 2017). Lophiostomataceae Sacc. Notes: The family was introduced as ‘‘Lophiostomeae’’ by Nitschke (1869) and recorded as ‘‘Lophiostomaceae’’ in 123 42 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 25 Hermatomyces bauhiniae (MFLU 18-1381, holotype). a– c Sporodochia on natural substrate. d Vertical section through sporodochia. e–g Subicular hyphae. h Cylindrical conidia and lenticular conidia on host substrate. i, j Cylindrical conidia. k–m Mature lenticular conidia. n Conidia. o Germinated conidium. Scale bars: a = 1 cm, b = 500 lm, c, d, f = 100 lm, o = 50 lm, h– m = 20 lm, e, g = 10 lm Saccardo (1883), but was accepted as Lopiostomataceae in Mugambi and Huhndorf (2009). Members of this family are saprobes and are found in terrestrial, aquatic and marine habitats (Tanaka and Harada 2003; Zhang et al. 2009c; Hirayama and Tanaka 2011; Hyde et al. 2013; Thambugala et al. 2015; Jones et al. 2015; Tennakoon et al. 2018). A generic re-evaluation of Lophiostomataceae by Thambugala et al. (2015) based on multi-locus phylogenies using LSU, SSU, ITS and TEF1-a sequence data, integrated 16 genera into this family, and resolved the phylogenetic and morphological boundaries of Lophiostomataceae. Hashimoto et al. (2018) revised the taxonomical classification of Lophiostoma bipolare (K.D. Hyde) E.C.Y. Liew, Aptroot & K.D. Hyde, and proposed a novel genus Lentistoma A. Hashim., K. Hiray. & Kaz. Tanaka to accommodate Lophiostoma sensu stricto. Four additional new genera and two new genera combinations were recognized from the other lineages of the L. bipolare complex. Therefore, there are currently 23 genera in Lophiostomataceae. Vaginatispora K.D. Hyde Notes: Vaginatispora was introduced by Hyde (1995b) with V. aquatica K.D. Hyde as the type species. Vaginatispora aquatica was initially placed in Massarinaceae, but Zhang et al. (2014a) revealed its phylogenetic affinity with Lophiostomataceae. Subsequently, Thambugala et al. (2015) considered Vaginatispora as a separate genus within Lophiostomataceae based on multi-gene phylogeny and morphology, of which V. aquatica and V. fucklii were included. Thereafter, V. appendiculata (Wanasinghe et al. 2016), V. armatispora (: Massarina armatispora K.D. Hyde, Vrijmoed, Chinnarij & E.B.G. Jones) (Hyde et al. 1992; Liew et al. 2002; Wanasinghe et al. 2016), V. microarmatispora (Devadatha et al. 2017), V. amygdali and V. scabrispora (Hashimoto et al. 2018) were introduced based on morphology and phylogeny. However, V. fuckelii was transferred to a new genus Neovaginatispora (Hashimoto et al. 2018) based on phylogenetic distinction and its typical characters of erumpent ascomatal and a thinner peridium. During observations of microfungi from marine (mangrove) and freshwater habitats, Vaginatispora palmae sp. nov. and V. armatispora were reported and illustrated herein. An updated phylogenetic tree is 43 provided, which includes all Vaginatispora taxa and closely related genera in Lophiostomataceae. Vaginatispora palmae S.N. Zhang, J.K. Liu & K.D. Hyde, sp. nov. Index Fungorum number: IF556316; Facesoffungi number: FoF05089; Fig. 26 Etymology: The epithet reflects the family of host plant. Holotype: MFLU 18-1586 Saprobic on immersed rachis of Nypa fruticans. Sexual morph Ascomata in vertical section 250–340 lm high, 215–385 lm diam. ( x = 310 9 326 lm, n = 10), dark brown to black, scattered, semi-immersed, erumpent, subglobose to elongated, base flatted, coriaceous to carbonaceous. Ostiole crest-like, variable in shape, central papillate. Peridium 15–38 lm wide, wider at the apex and thinner at the base, composed of several pale brown to brown cells of textura angularis, cells towards the inside lighter, at the outside darker, somewhat flattened, fusing and with the host tissues. Pseudoparaphyses 1–2.5 lm wide, hypha-like, numerous, septate, rarely branched and anastomosed, tapering towards the apex. Asci 89– 115 9 12–20 lm ( x = 100.5 9 16.0 lm, n = 20), 8spored, bitunicate, fissitunicate, cylindric-clavate, with a short bulbous pedicel, rounded at the apex, with an ocular chamber. Ascospores 23–45 9 6–9 lm, ( x = 35.3 9 7.5 lm, n = 30), hyaline, uniseriate or overlapping to biseriate, 1-septate, occasionally producing pseudosepta, slightly constricted at the central septum, cell above central septum swollen, guttulate, smooth-walled, surrounded by a narrow mucilaginous sheath and drawn out towards each end to form tapering appendages, 6–8 lm long. Asexual morph Undetermined. Culture characteristics: Ascospores germinating on PDA within 24 h. Colonies growing on PDA reaching 2 cm diam. after 21 days at 25 C, the off-white hyphae in first week, then becoming grayish blue and dark bluish, composed of brown to dark brown, septate, smooth or verrucose hyphae. Material examined: THAILAND, Ranong Province, on immersed rachis of Nypa fruticans Wurmb (Arecaceae), 3 December 2016, S.N. Zhang, SNT92 (MFLU 18-1586, holotype), ex-type living culture (MFLUCC 18-1526); ibid. (HKAS 102207, isotype). GenBank numbers: ITS: MK085055, LSU: MK085059, SSU: MK085057, TEF1-a: MK087657. Notes: Most species of Vaginatispora are found in tropical regions and commonly occur in freshwater and marine environments, but only one species V. nypae Jayasiri, E.B.G. Jones & K.D. Hyde (Jayasiri et al. 2019) has recently been recorded from a Nypa fruticans. Almost all Vaginatispora species have 1-septate ascospores with terminal appendages or sheaths (Hashimoto et al. 2018). 123 44 Fig. 26 Vaginatispora palmae (MFLU 18-1586, holotype). a, b Appearance of stromata on host surface. c Structure of peridium. d Vertical section through the stromata with ascomata. e Pseudoparaphyses. f–h Ascus. i–k Ascospores. l Ascospore in India Ink, with 123 Fungal Diversity (2019) 96:1–242 clear terminal appendages and narrow sheath. Sacle bars: b = 500 lm, a = 200 lm, d = 100 lm, c, e, f–h = 20 lm, i– l = 10 lm Fungal Diversity (2019) 96:1–242 45 Detailed characters of ascospores are useful for distinguishing taxa at the species level, and molecular sequence data is key for identification of taxa in this group. The multi-gene analysis includes all species in this genus and the new isolate Vaginatispora palmae clustered with V. amygdali (Fig. 28). Vaginatispora palmae differs from V. amygdali because the latter species has a lateral pad-like structure within the sheath and an internal chamber at both ends of the ascospores, while V. palmae lacks those structures. They also differ in peridium width (15–38 lm vs. 37.5–62.5 lm), ascus (mean: 100.5 9 16 lm vs. 115 9 18.5 lm) and ascospores size (mean: 35.3 9 7.5 lm vs. 30.6 9 8.8 lm). In addition, polymorphic nucleotide comparison shows that these two strains differ in eight positions in the ITS1 and ITS2 regions including two gaps, and differ in eight positions in the TEF1-a region. The difference in molecular sequence data also distinguish V. palmae from V. amygdali, therefore, we introduce a new species V. palmae. site1–4–1, (MFLU 18-1229), living culture (MFLUCC 18-0247). GenBank numbers: ITS: MK085056, LSU: MK085060, RPB2: MK087669, SSU: MK085058, TEF1-a: MK087658. Notes: The strain MFLUCC 18-0247 that was isolated and described in this study, phylogenetically clustered together with Vaginatispora armatispora (HKUCC 1562) (Wanasinghe et al. 2016) with high support (Fig. 28). On the other hand, these two V. armatispora strains (MFLUCC 18-0247 and HKUCC 1562) have only one base pair difference in the ITS1 region. These two isolates are quite similar, but isolate MFLUCC 18-0247 has a thicker peridium than that of HKUCC 1562. Based on such insignificant differences in morphology and molecular data, we consider the isolate (MFLUCC 18-0247) and V. armatispora are conspecific. Vaginatispora armatispora can be distinguished from other members in Vaginatispora because it has pale brown ascospores. Vaginatispora armatispora (K.D. Hyde, Vrijmoed, Chinnaraj & E.B.G. Jones) Wanas., E.B.G. Jones & K.D. Hyde Facesoffungi number: FoF05060; Fig. 27 Saprobic on decaying wood, submerged wood in freshwater and marine habatits. Sexual morph Ascomata 325–370 lm diam. 300–340 lm high, ( x = 350 9 317 lm, n = 10), scattered, usally immersed or semi-immersed, globose to subglobose, black. Ostiole black, crest-like, central, filled with brown cells. Peridium 20–32 lm wide, 2–3 layers, composed of dark brown to hyaline, thick-walled cells of textura porrecta at side and textura angularis at the base. Hamathecium comprising 2– 3 lm wide, septate, hyaline, pseudoparaphyses composing situated between and above the asci, embedded in a gelatinous matrix. Asci 94–118 9 15–17 lm ( x = 106 9 16 lm, n = 20), 8-spored, bitunicate, fissitunicate, cylindrical-clavate, short pedicellate, apically rounded with an ocular chamber, hyaline to subhyaline. Ascospores 29–34 9 7–9 lm ( x = 32 9 8 lm, n = 30), fusiform, hyaline to pale brown, smooth-walled, 1-septate, constricted at the septum, tapering to pointed apices, distinct large guttules in each cells, with distinct hyaline appendages, surrounded by a narrow mucilaginous sheath or not. Asexual morph undetermined. Culture characteristics: Ascospores germinating on PDA within 24 h. Colonies on MEA at room temperature reaching 3 cm diam. in three weeks, mycelium pale brown at first, grayish brown after 3 weeks, composed pale brown to brown, septate, smooth hyphae. Material examined: THAILAND, Prachuap Khiri Khan Province, Bang Saphan District, Sai Khu Waterfall, on submerged decaying wood, August 2017, V. Kumar, Massariaceae Nitschke Notes: The family Massariaceae was introduced by Nitschke (1869) to accommodate the genus Massaria De Not. with M. inquinans (Tode) De Not. as the type species. The family is characterized by immersed, globose, subglobose to pyriform, ascomata, with a thin-walled peridium comprising cells of textura angularis, oblong to cylindrical asci with a wide ocular chamber and refractive ring, and large, oblong to ellipsoidal ascospores surrounded by a gelatinous sheath (Voglmayr and Jaklitsch 2011; Hyde et al. 2013, 2016). Based on combined LSU, SSU and TEF1-a sequence data, Hyde et al. (2016) introduced the genus Neomassaria Mapook, Camporesi & K.D. Hyde in Massariaceae but Ariyawansa et al. (2018) introduced Neomassariaceae Ariyawansa, Jaklitsch & Voglmayr to accommodate this genus. Based on morphology and LSU phylogeny, Huanraluek et al. (2018) described the genus Massarioramusculicola Huanraluek., Thambugala & K.D. Hyde. At present, the family comprises 31 Massaria species and a single Massarioramusculicola species. In this study, we introduced a new genus Paramassaria typified with Paramassaria samaneae. Paramassaria Samarak., & K.D. Hyde, gen. nov. Index Fungorum number: IF555521; Facesoffungi number: FoF05213 Etymology: Named because of its morphological similarity to Massaria. Saprobic on branches. Sexual morph Ascomata immersed or semi-immersed in the yellow stained host, solitary, scattered, coriaceous, globose to subglobose, brown to dark brown. Ostiole central, raised, flattened around the dome-shaped ostiole, surrounded by blackish stromatic zone on the host, minutely papillate. Peridium 123 46 Fig. 27 Vaginatispora armatispora (MFLU 18-1229, new record). a Ascomata on submerged wood. b Section of ascoma. c–e Section of peridium. f Pseudoparaphyses. g–j Asci. k–m Ascospores. 123 Fungal Diversity (2019) 96:1–242 n Germinating ascospore. o, p Culture on PDA. Scale bars: b = 100 lm, c–e, g–n = 20 lm, f = 10 lm Fungal Diversity (2019) 96:1–242 47 Fig. 28 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU, SSU, TEF1-a and RPB2 sequence data for Vaginatispora species and several closely related genera in Lophiostomataceae. Related sequences are taken from Thambugala et al. (2015), Wanasinghe et al. (2016), Devadatha et al. (2017) and Hashimoto et al. (2018). Thirty-four strains are included in the combined analyses which comprise 4243 characters (605 characters for ITS, 832 characters for LSU, 895 characters for SSU, 894 characters for TEF1-a, 1017 characters for RPB2) after alignment. Angustimassarina populi (MFLUCC 13–0034) and A. acerina (MFLUCC 14-0505) (Amorosiaceae, Pleosporales) are used as the outgroup taxa. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of –17911.101212 is presented. The matrix had 1059 distinct alignment patterns, with 20.20% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.248793, C = 0.250421, G = 0.268979, T = 0.231807; substitution rates AC = 1.458334, AG = 3.325243, AT = 1.282598, CG = 1.395939, CT = 8.065988, GT = 1.000000; gamma distribution shape parameter a = 0.592418. Maximum parsimony analysis of 3144 constant characters and 854 informative characters resulted in two equally most parsimonious tree with a tree length as 2508 (CI = 0.622, RI = 0.822, RC = 0.511, HI = 0.378). Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 75 are placed above and below the branches respectively. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are in bold. The ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue wide around the ostiole, outer layer; thick-walled, reddish to dark brown cells of textura angularis, inner layer; thinwalled hyaline cells of textura angularis. Hamathecium comprising wide, hyaline, cylindrical to filiform, septate, branched pseudoparaphyses. Asci 8-spored, bitunicate, oblong to cylindrical, absence of prominent ocular chamber, shortly pedicellate. Ascospores overlapping 1–2-seriate, ellipsoid to broadly fusiform, initially hyaline, 1-septate becoming brown and 3-septate, constricted at the septum, surrounded by a hyaline gelatinous sheath at maturity. Asexual morph Undetermined. Notes: Paramassaria samaneae was collected on a dead branch of Samanea saman from Thailand. This taxon is similar to Massaria in its ascomata, asci and ascospore characters. The wide ostiole and absence of an ocular chamber in the asci distinguishes our taxon from Massaria. However, the combined phylogenetic analyses (Fig. 29) separate our taxon as a basal clade to Massaria. The closest BLAST search with the LSU and TEF1-a sequences of MFLU 17-1551 were GQ221892 (with 93% identity) Rhytidhysteron opuntiae and FJ161103 (with 93% identity) Gloniopsis praelonga belonging to Hysteriaceae. Only ITS, 123 48 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 29 Phylogram generated from Bayesian analysis based on combined LSU, SSU and TEF1-a sequence data for Massariaceae and several closely related families in Pleosporales. Related sequences are taken from Voglmayr and Jaklitsch (2011), Hyde et al. (2016) and Ariyawansa et al. (2018). One-hundred and one strains are included in the combined analyses which comprise 2519 characters (806 characters from LSU, 840 characters for SSU, 873 characters for TEF1-a) after alignment. Bayesian posterior probabilities (BYPP) equal to or greater than 0.9 are on branches. The newly generated sequences are indicated in bold and blue LSU and SSU sequences are available for Massarioramusculicola and resulted in inconsistent topologies with the absence sequence data. In addition, we observed that our strains have probabilities to cluster in Aigialacaeae as a basal clade. However, Massarioramusculicola differs from Massaria and Paramassaria by having a peridium with two strata with dark brown to hyaline cells of textura angularis, rarely branched, septate and trabeculate pseudoparaphyses and 3-septate hyaline ascospores. Including parallel morphology to Massaria, we prefer to place our strains into a new genus in Massariaceae. Voglmayr and Jaklitsch (2011) suggested that the ascospore and pseudothecia characters are important for the identification of phylogenetically distinct Massaria species. The strong host-specificity among Massaria species makes it complex for accurate identification (Voglmayr and Jaklitsch 2011). It might also not be a surprise if future discoveries of more species within Massariaceae converge the clade and there is a need to aggregate them into one genus, but it is unwise to do this at this stage. Type species: Paramassaria samaneae Samarak., & K.D. Hyde Paramassaria samaneae Samarak & K.D. Hyde, sp. nov. Index Fungorum number: IF555522; Facesoffungi number: FoF05214; Fig. 30 Etymology: Name based on the host Samanea, from which it was collected. Holotype: MFLU 17-1551 Saprobic on a dead branch of Samanea saman. Sexual morph Ascomata 520–580 lm high ( x = 540 lm, n = 15), 410–560 lm diam. ( x = 490 lm, n = 15), immersed or semi-immersed in the yellow stained host, solitary, scattered, coriaceous, globose to subglobose, brown to dark brown. Ostiole central, flattened on the top, surrounded by blackish stromatic zone on the host, ostiolar canal with hyaline periphyses. Peridium 20–43 lm wide ( x = 33.5 lm, n = 25), inner cell layer hyaline, thin and outer cell layer comprising dark brown textura angularis cells. Hamathecium comprising 4.2–10 lm wide ( x = 6.6 lm, n = 25), hyaline, cylindrical to filiform, septate, branched pseudoparaphyses. Asci 190–215 9 45.5– 56 lm ( x = 212.5 9 50.5 lm, n = 20), 8-spored, 49 bitunicate, oblong to cylindrical, short pedicellate. Ascospores 52.5–71.5 9 13.5–18.5 lm ( x = 61.5 9 16 lm, n = 50), l/w 3.9, overlapping 1–2-seriate, brown, ellipsoid to broadly fusiform, 3-septate, constricted at the septa, with 4 guttlues, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph Undetermined. Material examined: THAILAND, Chiang Rai Province, Muang District, Mae Fah Luang University, on a dead branch of Samanea saman (Fabaceae), 2 August 2017, MC. Samarakoon, SAMC002 (MFLU 17-1551, holotype; HKAS 102338, isotype). GenBank numbers: LSU: MK108190, MK108191, SSU: MK108187, MK108188, TEF1-a: MK105747, MK105747. Phaeosphaeriaceae M.E. Barr Notes: Phaeosphaeriaceae is the largest family in Pleosporales and was introduced by Barr (1979). It is typified with Phaeosphaeria with the type species Phaeosphaeria oryzae (Phookamsak et al. 2014). Phaeosphaeriaceae comprises 52 genera (Wijayawardene et al. 2018a). Phaeosphaeriaceae is a heterogenous family of saprobes, pathogens and endophytes on plants (Tennakoon et al. 2016a; Phookamsak et al. 2017; Tibpromma et al. 2018). The family is characterized by bitunicate asci with hyaline, yellow or brown spores that are aseptate or septate and holoblastic or enteroblastic conidiogenesis. Seven new species belonging to Muriphaeosphaeria (M. angustifoliae), Neosetophoma (N. miscanthi, N. salicis), Phaeopoacea (P. asparagicola), Phaeosphaeria (P. penniseti) and Nodulosphaeria (N. thalictri, N. aquilegiae) are introduced. Muriphaeosphaeria Phukhamsakda et al. Notes: The genus Muriphaeosphaeria was introduced by Phukhamsakda et al. (2015) based on the type species Muriphaeosphaeria galatellae found on dead and drying stems of Galatella villosa (L.) Rchb.f. (Asteraceae) in Russia. The morphological characters of the genus comprise superficial ascomata with a peridium containing thinwalled brown cells of textura angularis, cellular pseudoparaphyses and muriform ascospores. The asexual morph is coelomycetous and is characterized by conidiomata with a thick, hyaline, inner wall layer and cylindrical to subclavate, 1–3-transversely septate, brown conidia (Phukhamsakda et al. 2015). Muriphaeosphaeria contains three species (Index Fungorum 2019). Cultures and sequences are available. Muriphaeosphaeria angustifoliae D. Pem, Gafforov & K.D. Hyde, sp. nov. Index Fungorum number: IF555432; Facesoffungi number: FoF05109; Fig. 31 123 50 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 30 Paramassaria samaneae (MFLU 17–1551, holotype). a, b Ascomata on the substrate. c Horizontal section of ascoma. d Vertical section of ascoma. e Ostiole section. f Peridium (in water). g Pseudoparaphyses. h–j Asci. k–s Ascospores (s-in Indian ink). Scale bars: a = 1000 lm, b–d = 200 lm, e = 100 lm, h–j = 50 lm, f, g, k–s = 20 lm Etymology: Name reflects the host from which the fungus was isolated. Holotype: TASM 6138 Saprobic on dead branches of Perovskia angustifolia. Sexual morph Ascomata 200–600 lm high 9 200– 500 lm diam. ( x = 228.2 9 226.5 lm, n = 10), scattered or sometimes clustered, or solitary, semi-immersed to superficial, globose to subglobose, brown to dark brown, with papillate central ostiole. Ostiole 17–18 lm high 9 28–33 lm wide, slightly raised, centrally located, lacking periphyses. Peridium 70–90 lm thick, smooth-walled comprising 2–3 layers of brown to dark brown, pseudoparenchymatous cells of textura angularis. Hamathecium comprising numerous, dense, 1–1.5 lm wide, filiform, broadly cellular pseudoparaphyses, branched and anastomosing, embedded in gelatinous matrix. Asci 125– 175 9 13–15 lm ( x = 148 9 14.5 lm, n = 10), 8-spored, bitunicate, fissitunicate, cylindrical to cylindric-clavate, short pedicellate or subsessile, apically rounded with an ocular chamber. Ascospores 21–28 9 9–11 lm ( x = 25.9 9 10.1 lm, n = 10), overlapping uniseriate, initially hyaline, becoming yellowish brown at maturity, broadly fusiform, multiseptate, constricted at the middle septa, narrowly rounded at both ends, smooth-walled, swollen, lacking a mucilaginous sheath. Asexual morph Undetermined. Culture characteristics: Ascospores germinating on MEA within 48 h. Colonies growing on MEA, reaching 3 cm diam. in 1 week at 16 C. Mycelium superficial, surface smooth, circular, hairy, margin well defined, shiny, from above white, reverse pale yellow and grey in the middle. Material examined: UZBEKISTAN, Surxondaryo, Boysun, Omonxona Village, South-Western Hissar Mountains, on dead branches of Perovskia angustifolia (Lamiaceae), 6 April 2016, Y. Gafforov, YG-S29-2 (TASM 6138, holotype; MFLU 17-0070, isotype), ex-type living culture (MFLUCC 18-1124). GenBank numbers: ITS: MK101002, LSU: MK100999, SSU: MK100997. Notes: Muriphaeosphaeria angustifoliae constitutes a strongly supported independent lineage closely nested with M. ambrosiae (MFLU 15-1971). Muriphaeosphaeria ambrosiae is only known from its asexual morph and cannot be compared to M. angustifoliae. 51 Muriphaeosphaeria angustifoliae is morphologically more similar to M. galatellae (Phukhamsakda et al. 2015) in having superficial ascomata and broadly fusiform multiseptate ascospores lacking a mucilaginous sheath. However, M. angustifoliae has larger ascomata (200–600 lm 9 200–500 lm v/s 114–180 9 167–263 lm), a wider peridium (70–90 lm v/s 12–26 lm), narrower hamathecium (1–1.5 lm v/s 2–3 lm) and longer asci (125–175 9 13–15 lm v/s 53–86 9 9–17 lm). We therefore, introduce M. angustifoliae as a new species in Phaeosphaeriaceae. Neosetophoma Gruyter et al. Notes: Neosetophoma was introduced by de Gruyter et al. (2010) and is typified by N. samararum (Desm.) Gruyter, Aveskamp. & Verkley. Neosetophoma comprises 15 species (Wijayawardene et al. 2017a) and contains important plant pathogens, saprobes and soil fungi (Phookamsak et al. 2014; Karunarathna et al. 2017). The genus is characterized by globose to irregular conidiomata with papillate ostioles and yellowish conidia attenuated at one end (Wijayawardene et al. 2017a). In this study Neosetophoma miscanthi and Neosetophoma salicis are introduced. Neosetophoma miscanthi Karun., C.H. Kuo & K.D. Hyde, sp. nov. Index Fungorum number: IF556266; Facesoffungi number: FoF05848; Fig. 32 Etymology: Name reflects the host genus, Miscanthus from which the fungus was isolated. Holotype: MFLU 18-2675 Saprobic on dead stems of Miscanthus gigantius (Poaceae), as raised, black, shiny globular structures on the host surface. Sexual morph Ascomata 90–130 lm high, 110– 120 lm diam. ( x = 109.5 9 115 lm, n = 5), solitary, gregarious, immersed under epidermis, becoming erumpent through host surface, subglobose, or polygonal at sides, uni-loculate, glabrous, ostiolate, with minute papilla; papilla 30–50 lm long (n = 5), carbonaceous, easily broken. Peridium 19–20 lm wide, thin-walled, of equal thickness, composed of few layers of dark brown to brown, pseudoparenchymatous cells of textura angularis. Hamathecium of dense, filamentous, septate, cellular pseudoparaphyses, embedded in a hyaline gelatinous matrix. Asci 45–50 9 10–12 lm ( x = 48 9 10.5 lm, n = 20), 8spored, bitunicate, fissitunicate, cylindric-clavate to obclavate, slightly broad at the base with short rounded pedicel, apically rounded. Ascospores 18–21 9 4.7– 5.3 lm ( x = 20 9 5 lm, n = 40), overlapping 1–2-seriate, hyaline, fusiform, with acute ends, occasionally 3-septate, cell near the septum slightly larger, smooth-walled, guttulate. Asexual morph Undetermined. 123 52 Fungal Diversity (2019) 96:1–242 Fig. 31 Muriphaeosphaeria angustifoliae (TASM 6138, holotype). a, b Appearance of ascomata on host surface. c Section through an ascoma. d Peridium. e Hamathecium. f. Ostiole. g-i Asci. j-l Ascospores. m Ascospore germination on MEA after 48 h. n, o Culture characters on MEA (n from above view, o from below view). Scale bars: a = 2000 lm, b = 500 lm, c = 100 lm, g–i = 50 lm, d, f, m = 20 lm, e, j–l = 10 lm Culture characteristics: Ascospores germinating on PDA within 12 h reaching 20 mm diam. after 7 d at 25 C, circular, floccose, flat with even margin, initially pale yellow becoming greenish grey; reverse yellowish grey. 123 Fungal Diversity (2019) 96:1–242 53 Fig. 32 Neosetophoma miscanthi (MFU 18-2675, holotype). a, b Appearance of ascomata on Miscanthus gigantius. c. Section through ascoma. d. Section through ostiole. e. Section through peridium. f. Pseudoparaphyses. g–j. Different developing stages of the asci. l–n. Ascospores. o. Germinating ascospore. Scale bars: c–e = 50 lm, f = 20 lm, g–o = 10 lm Material examined: TAIWAN, Chia Yi Province, Kwang Hwa, Miscanthus gigantius (Poaceae), 18 March 2018, A. Karunarathna AKTW 31 (MFLU 18-2675, holotype), ex-type living culture (FU31023). GenBank numbers: ITS: MK503820, LSU: MK503826, SSU: MK503832. Notes: Neosetophoma miscanthi is morphologically similar to N. guiyangensis but is distinct in its asci and ascospore characters. In N. miscanthi asci are cylindricclavate to obclavate, broad at the base whereas the asci of N. guiyangensis are cylindrical to cylindric-clavate. Furthermore, N. miscanthi generally has smaller ascomata, peridia and asci parameters than N. guiyangensis (Hyde et al. 2018a). Neosetophoma miscanthi prominently consists with 3-septate ascospores, while N. guiyangensis prominently consisting with 1–3(–5)-septate ascospores. Neosetophoma miscanthi was found on Poaceae while N. guiyangensis was found on tree branches. In phylogenetic analysis the N. miscanthi and N. guiyangensis separates with high support (90% ML/0.99 BYPP) Associated with twigs and branches. Sexual morph Undetermined. Asexual morph Conidiomata 8–10 9 3– 4.5 lm, immersed in host tissue, scattered, erumpent, discoid, circular, with locules, ostiolate. Ostioles at the same level, with flattened top. Peridium brown to pale brown, comprising a few layers of cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells lining the inner cavity, hyaline, smooth, phialidic with prominent periclinal thickening and thick channel, globose to doliiform, 8–10 9 3–4.5 lm, formed from the inner most layer of pycnidial wall, hyaline, smooth-walled. Conidia (6.9–)8–10 9 2–2.8(–3.1) lm ( x = 8.7 9 2.8 lm, n = 30), unicellular, allantoid, hyaline, smooth-walled with thick layer. Material examined: UZBEKISTAN: Surxondaryo Province, Boysun District, Qizilnaur Village, South-Western Hissar Mountains, twigs and branches of Salix sp. (Salicaceae), 17 June 2016, Y. Gafforov, YG-S109-2 (TASM 6147, holotype; MFLU 17-0118, isotype), ex-type living culture (MFLUCC). GenBank numbers: ITS: MK608025, LSU: MK608026. Notes: Neosetophoma salicis is introduced as a new species based on phylogenetic analysis, as it forms a separate branched as a sister taxon to N. rosae Jayasiri et al., which was isolated from twigs and branches of Rosa canina L. in Italy (isolate MFLUCC 17-0844, Wanasinghe Neosetophoma salicis Norphanphoun, Gafforov, T.C. Wen & K.D. Hyde, sp. nov. Index Fungorum number: IF555517; Facesoffungi number: FoF06000; Fig. 33 Etymology: The specific epithet reflects the host genus. Holotype: TASM 6147 123 54 Fungal Diversity (2019) 96:1–242 Fig. 33 Neosetophoma salicis (TASM 6147, holotype). a Stromatal habit in wood. b Fruiting bodies on host surface. c Surface of fruiting bodies showing the black ostioles. d Cross section of conidioma. e Peridium. f Ostiolar. g–h Conidiogenous cell containing conidia. i Conidia. Scale bars: d = 100 lm, e, f = 50 lm, g, h = 10 lm, i = 5 lm et al. 2018a) with strong support (Fig. 38). However, they are different in morphological characteristics, the conidia of N. salicis are shorter than N. rosae (8.7 9 2.8 versus 11.9 9 2.5 lm) and conidia lack septa. Shoemaker (1976) re-circumscribed Nodulosphaeria and transferred various species to Ophiobolus. Later, many species were added to the genus based on morphology (Shoemaker 1984; Shoemaker and Babcock 1987; Zhang et al. 2012). There are 71 epithets listed under Nodulosphaeria (Index Fungorum 2019). Mapook et al. (2016) revised the genus and confirmed its phylogenetic placement in Phaeosphaeriaceae. Nodulosphaeria Rabenh., Klotzschii Herb. Notes: The genus Nodulosphaeria was introduced by Rabenhorst (1858) with N. hirta Rabenh. as type species. It is characterized by brown setae at the ascomata apex, and three- to multi-septate ascospores, with a swollen cell and some with terminal appendages (Holm 1961; Shoemaker 1984). The genus was re-examined by Zhang et al. (2012) and Phookamsak et al. (2014) based on type material of N. hirta. Nodulosphaeria was previously treated as a synonym of Leptosphaeria, but later reinstated by Holm (1957). Nodulosphaeria was accommodated in Phaeosphaeriaceae by Barr (1979) based on morphology and similarity to Ophiobolus as the ascospores form enlarged cell(s). 123 Nodulosphaeria aquilegiae Chaiwan, Camporesi & K.D. Hyde, sp. nov. Index Fungorum number: IF555508; Facesoffungi number: FoF05200; Fig. 34 Fig. 34 Nodulosphaeria aquilegiae (MFLU 17-1273, holotype). a, c, c d Ascomata on a dead stem of Aquilegia viscosa. b, i Peridium. e, f Section of ascomata. g Pseudoparaphyses. h-n Asci. o-r Ascospores. Scale bars: c, d = 200 lm, b, e, f–i = 50 lm, j–n = 20 lm, o– r = 10 lm Fungal Diversity (2019) 96:1–242 55 123 56 Etymology: Name reflects the host from which the fungus was isolated. Holotype: MFLU 17-1273 Saprobic on dead stem of Aquilegia viscosa. Sexual morph Ascomata 150–195 lm high 9 95–195 lm diam. ( x = 167.8 9 142.2 lm, n = 5), superficial, solitary, scattered, uniloculate, globose to subglobose, dark brown to black, conspicuous on the host. Ostioles 50–60 9 45– 60 lm ( x = 53.5 9 52.5 lm, n = 5), protruding from the centre of the ascomata, without setae. Peridium 40–80 lm wide, comprising several layers of brown to dark brown cells of textura angularis, outer layer darkened cells and inner layer pale brown to hyaline cells. Hamathecium comprising numerous, filiform, septate, hyaline pseudoparaphyses. Asci 70–100 9 7–11 lm ( x = 83 9 9 lm, n = 20), 8-spored, bitunicate, short pedicellate with clubshaped pedicel, or sessile, apically rounded. Ascospores 20–30 9 4.5–5.5 lm ( x = 25 9 5 lm, n = 20), overlapping 1–2-seriate, initially hyaline, becoming yellowish brown at maturity, fusiform, ellipsoidal to subcylindrical, slightly curved, upper part shorter and wider than the lower part, 4-transversely septate, constricted at the middle septum, ends remaining lighter and cone-shaped, with narrowly rounded ends, wall firm, smooth-walled, sometimes guttulate, with appendages (3–5 lm long, 3–4 lm wide) at both ends. Asexual morph: undetermined. Material examined: ITALY, Province of Forlı̀-Cesena [FC], Passo la Calla - Santa Sofia, on a dead aerial stem of Aquilegia viscosa Gouan (Ranunculaceae), 14 July 2017, Erio Camporesi, IT3398 (MFLU 17-1273, holotype). GenBank numbers: ITS: MK579853, LSU: MK579854, SSU: MK579855, TEF1-a: MK579856. Notes: Nodulosphaeria aquilegiae was found on dead, aerial stems of Aquilegia viscosa in Italy. Ascomatal characteristics resemble N. multiseptata and N. sambuci in having superficial, solitary, uniloculate, globose to subglobose ascomata (Tibpromma et al. 2017). This is the first record of a Nodulosphaeria species on Aquilegia viscosa. Nodulosphaeria thalictri D. Pem, Camporesi & K.D. Hyde, sp. nov. Index Fungorum number: IF555433; Facesoffungi number: FoF05110; Fig. 35 Etymology: Name reflects the host from which the fungus was isolated. Holotype: MFLU 17-1241 Saprobic on dead stems of Thalictrum sp. Sexual morph Ascomata 217–305 lm diam. 9 185–304 lm high ( x = 263.7 9 217.9 lm, n = 10), immersed or semi-immersed, solitary or scattered, subglobose to obpyriform, dark brown to black, papillate. Ostiole 70–75 lm diam. ( x = 71.2 lm, n = 10), papillate, protruding from substratum with numerous internal brown to dark brown cells of 123 Fungal Diversity (2019) 96:1–242 textura globulosa. Peridium 12–16 lm wide, comprising 4–5 layers of dark brown cells of textura angularis, flattened at the inside. Hamathecium comprising 1.5–2.5 lm wide, filiform, septate, branching pseudoparaphyses. Asci 77–100 9 7–12 lm ( x = 84.1 9 10.1 lm, n = 10), 8spored, bitunicate, cylindric-clavate, slightly curved, short pedicellate, apically rounded, with an ocular chamber. Ascospores 29–35 9 4–5 lm ( x = 32.3 9 4.5 lm, n = 20), overlapping 3–4 seriate, hyaline when immature, becoming pale yellowish to greenish at maturity, long fusiform, 7–8 septate, straight or slightly curved, thick and smooth-walled, with terminal appendages. Asexual morph Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at 16 C and germ tubes produced from middle septa. Colonies growing on MEA, reaching 1 cm diam. in 1 week. Mycelium superficial, medium sparse, irregular, flat, smooth, effuse, slightly irregular margin, from above whitish, reverse yellowish to pale brown. Material examined: ITALY, Province of Forlı̀-Cesena [FC], Lago Pontini – Bagno di Romagna, on dead aerial stems of Thalictrum sp. (Ranunculaceae), 5 July 2017, E. Camporesi (MFLU 17-1241, holotype; HKAS 97500, isotype), ex-type living culture (MFLUCC 18-1138). GenBank numbers: ITS: MK 101000, LSU: MK 100998, SSU: MK 101001, TEF1-a: MK128971. Notes: Nodulosphaeria thalictri is similar to N. hirta (type species of Nodulosphaeria) in having cylindric-clavate, slightly curved, short pedicellate asci and a peridium with dark brown cells of textura angularis, however N. thalictri has narrower pseudoparaphyses (1.5–2.5 lm v/s 2.5–3.5 lm), shorter and wider ascospores ( x = 32.3 9 4.5 lm v/s x = 55 9 5 lm) and the presence of terminal appendages (Phookamsak et al. 2014). Our phylogenetic analyses place N. thalictri close to N. aconiti (MFLUCC 13–0728). Nodulosphaeria thalictri is similar to N. aconiti in having a peridium composed of cells of textura angularis and branching pseudoparaphyses (Mapook et al. 2016). However, N. thalictri has longer asci (84 9 10 lm v/s 69 9 9 lm), more septa in the ascospores septation (7–8 v/s 4–septate), has long fusiform ascospores rather than cylindric-fusiform and possesses terminal appendages, which are lacking in N. aconiti. Nodulosphaeria thalictri is unique in having an ostiole with numerous internal brown to dark brown cells of textura globulosa, protruding externally, a feature not recorded in other Nodulosphaeria species. Phaeopoacea Thambug., Dissan. & K.D. Hyde Notes: Phaeopoacea was introduced by Thambugala et al. (2017b) to accommodate the Phaeopoacea species occurring as a saprobe on Poaceae (Hyde et al. 2017b; Thambugala et al. 2017b). Currently, the genus comprises Fungal Diversity (2019) 96:1–242 57 Fig. 35 Nodulosphaeria thalictri (MFLU 17-1241, holotype). a, b Appearance of ascomata on host surface. c Section through an ascoma. d Ostiole. e Peridium. f Hamathecium. g-i Asci. j-m Ascospores, arrow showing terminal appendages. n Germinated ascospore on MEA after 24 h. o, p Culture characters (o from above view, p from below view). Scale bars: a, b = 500 lm, c = 50 lm, d = 30 lm, g–i = 25 lm, e, n = 20 lm, f, j–m = 10 lm three species: P. festucae (type species), P. phragmiticola and P. muriformis. Sexual morphs of Phaepoacea are characterized by subepidermal ascomata with a central papillate ostiole, bitunicate asci and ascospores tapering towards the ends (Shoemaker and Babcock 1989; Hyde et al. 2017b). The asexual morphs from this genus have pycnidial conidiomata with brown conidia (Thambugala et al. 2017b). In this study, we introduce a novel species, Phaeopoacea asparagicola from Asparagaceae (Fig. 39). globose to compressed globose, coriaceous, dark brown to black, rough-walled, with short hyphae projecting from peridium, ostiolate. Ostiole centrally located, papillate, without periphysoids. Peridium 11–29 lm wide, composed of 4–7 layers of reddish brown to dark brown cells of textura angularis, inner layer composed of hyaline gelatinous cells. Hamathecium composed of numerous, dense, long, 2–7 lm wide ( x = 5 lm, n = 50), filiform, transversely septate, branched, anastomosing, cellular pseudoparaphyses. Asci 67–129 9 12–16 lm ( x = 98 9 14 lm, n = 30), 8-spored, bitunicate, fissitunicate, cylindric-clavate to broad cylindrical, with furcate pedicel, with ocular chamber visible when immature. Ascospores 22–33 9 6–10 lm ( x = 27 9 8 lm, n = 50), biseriate or overlapping, initially hyaline, becoming yellowish to brown at maturity, broad fusiform, narrow towards the apex, 3–7-transversely euseptate, constricted at the septa, with cell above central septum wider, smoothwalled, indentations present, surrounded by a 7–9 lm wide, mucilaginous sheath. Asexual morph Undetermined. Phaeopoacea asparagicola Phukhams., Akulov & K.D. Hyde, sp. nov. Index Fungorum number: IF555501; Facesoffungi number: FoF04826; Fig. 36 Etymology: Name refers to the host plant, Asparagus sp. Holotype: MFLU 18-1380 Saprobic on dead stems of Asparagus sp. Sexual morph Ascomata 200–365 9 110–314 lm ( x = 283 9 220 lm, n = 10), semi-immersed to immersed under epidermal tissue, solitary or sometimes gregarious, confluent, scattered, 123 58 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 36 Phaeopoacea asparagicola (MFLU 18-1380, holotype). a Appearance of ascomata on host surface. b Close up of ascomata. c, d Vertical sections of ascomata. e Section of partial peridium layer. f Cellular pseudoparaphyses. g–i Development state of asci. j–o Development state of ascospores. p Ascospores stained with india ink, sheath surrounding the entire ascospore. Scale bar: b = 500 lm, c, d = 100 lm, e–i = 50 lm, j–p = 10 lm Culture characteristics: Colonies on MEA, reaching 50 mm diam. after 4 weeks at 18 C. Culture olive-green radiating yellowish toward the edge, with dense mycelia, circular, flat, dull, fimbriate, radially furrowed, and slightly covered with white aerial mycelium; reverse black with radiating cream mycelium. Material examined: UKRAINE, Odessa region, Lyman district, Regional landscape park Tiligulskyt, on the overwintered stems of Asparagus sp. (Asparagaceae), 1 May 2014, A. Akulov, EX CWU (MYC) AS 5825 (MFLU 18-1380, holotype), ex-type living culture (MFLUCC 16-0379). GenBank numbers: ITS: MK443383, LSU: MK443379, RPB2: MK443387, SSU: MK443381, TEF1-a: MK443385. Notes: Phaeopoacea asparagicola is introduced as a new species based on morphological and phylogenetic evidence. The species formed a moderately-supported clade with P. festucae and P. phragmiticola (MLBS = 66; BYPP = 0.99). The morphology of P. asparagicola is similar to P. phragmiticola in having sub-epidermal, globose, papillate ascomata, cylindrical with short stalk asci, and conical at the apex with pale brown ascospores (Shoemaker and Babcock 1989; Hyde et al. 2017b). Besides these characters, P. asparagicola can be distinguished from P. phragmiticola and P. muriformis by having larger ascomata and asci and ascospores only have longitudinal septa. Phaeosphaeria I. Miyake Notes: Phaeosphaeria was introduced by Miyake (1909) with P. oryzae I. Miyake as the type species (Phookamsak et al. 2014). Previously, Phaeosphaeria was considered as a synonym of Leptosphaeria. However, based on both morphology and molecular analyses, it was considered that Phaeosphaeria differs from Leptosphaeria in having pseudoparenchymatous peridium, stagonospora-like asexual morphs and mostly monocotyledonous hosts while Leptosphaeria has scleroplectenchymatous peridium, phoma-like asexual morphs and mostly dicotyledonous hosts (Schoch et al. 2009; Zhang et al. 2012). Phaeosphaeria species are probably cosmopolitan in distribution, since they have been recorded from both temperate and tropical countries (i.e. Canada, China, Germany, Japan, Thailand, and Switzerland) (Shoemaker and Babcock 1989; Phookamsak et al. 2014, 2019). There are 211 epithets for Phaeosphaeria in 59 Index Fungorum (2019). We follow the latest treatment and updated account of Phaeosphaeria in Phookamsak et al. (2019). Phaeosphaeria penniseti is introduced as a novel species based on morphological and phylogenetic evidence (Figs. 37, 38). Phaeosphaeria penniseti Karun., C.H. Kuo & K.D. Hyde, sp. nov. Index Fungorum number: IF556267; Facesoffungi number: FoF05849; Fig. 37 Etymology: Name reflects the host genus, Pennisetum. Holotype: MFLU 18-2674 Saprobic on dead stems of Pennisetum purpureum (Poaceae). Sexual morph Ascomata 133–184 lm high 150– 243 lm diam. ( x = 158.5 9 196.5 lm, n = 5), black, immersed under host epidermis, visible as raised, to semiimmersed, solitary, subglobose to ampulliform, uni-loculate, ostiolate, with minute papilla. Peridium 18–20 lm wide at the sides, comprising 3–5 layers, of brown to dark brown pseudoparenchymatous cells, of textura angularis, hyaline to pale brown, flattened towards the inner layers. Hamathecium comprising numerous, filamentous, branched, septate pseudoparaphyses. Asci 55–70 9 10–15 lm ( x = 62.5 9 12.5 lm, n = 20), 8-spored, bitunicate, fissitunicate, cylindricclavate, pedicellate, rounded at the apex and with a shallow ocular chamber. Ascospores 26–29 9 4–5 lm ( x = 28 9 5 lm, n = 40), overlapping, 2-seriate, hyaline, becoming brown at maturity, narrowly fusiform, 5-septate, slightly wider at the second cell from above, guttulate at the beginning, rounded at the ends. Asexual morph: Undetermined. Material examined: TAIWAN, Chia Yi Province, Kwang Hwa, Pennisetum purpureum (Poaceae), 18 March 2018, A. Karunarathna AKTW 24 (MFLU 18-2674, holotype), ex-type living culture (FU31020). GenBank numbers: ITS: MK503819, LSU: MK503825, SSU: MK503831. Notes: In the NCBI BLAST search, Phaeosphaeria penniseti shows higher percentage identity to P. oryzae (99.80%). Phylogenetically, Phaeosphaeria penniseti forms a sister lineage with P. oryzae, the type species of the Phaeosphaeria with moderate support in maximum likelihood analysis (89% ML, Fig. 38). It differs from P. oryzae in having 5-septate ascospores, whereas, P. oryzae has 3-septate ascospores (Phookamsak et al. 2014). Pseudoberkleasmiaceae Phukhams & K.D. Hyde, fam. nov. Index Fungorum number: IF555489, Facesoffungi number: FoF05311 Saprobic on decaying wood. Sexual morph Undetermined. Asexual morph Hyphomycetous, dictyosporous. Colonies on natural substratum sporodochia, superficial, 123 60 Fungal Diversity (2019) 96:1–242 Fig. 37 Phaeosphaeria penniseti (MFLU 18-2674, holotype). a, b Appearance of ascomata on the host (Pennisetum purpureum, Poaceae). c Section through ascoma. d Section through ostiole. e Section through peridium. f Pseudoparaphyses (stained in cotton blue). g–i Different developing stages of the asci. j–m Ascospores. Scale bars: c, d = 50 lm, e = 25 lm, f–m = 10 lm compact, scattered, irregular, dark-brown to black, glistening. Mycelium immersed in the substrate, septate, branched. Conidiophores micronematous, mononematous, reduced, hyaline. Conidiogenous cells holoblastic, monoblastic, integrated, terminal, determinate. Conidia acrogenous, solitary, broadly ellipsoidal to obovoid, muriform, guttulate, smoothwalled, brown, olivaceous green, with or without guttules, usually with conidiogenous cell attached. Type genus: Pseudoberkleasmium pandanicola Tibpromma & K.D. Hyde, Fungal Diversity: 52 (2018) Notes: Pseudoberkleasmiaceae is introduced to accommodate a berkleasmium-like hyphomycete that is phylogenetically placed in the order Pleosporales. The family comprises Pseudoberkleasmium chiangmaiense and P. pandanicola, the latter as the generic type. Berkleasmium is polyphyletic, with species having similar characteristics, but which are phylogenetically distinct (Pinnoi et al. 2007; Hu et al. 2010; Lu et al. 2018b). A study by Tanney and Miller (2017) was able to obtained several fresh collections of Berkleasmium and compared those with the reference specimens of Berkleasmium concinnum. Re-examination of the type species placed the generic type of Berkleasmium in Tubeufiales. In this study, Pseudoberkleasmiaceae is Fig. 38 Phylogram generated from maximum likelihood analysis c based on combined ITS, LSU, SSU and TEF1-a sequence data for genera in Phaeosphaeriaceae. Related sequences are taken from Wanasinghe et al. (2018a) and Phookamsak et al. (2019). Onehundred and sixty-one strains are included in the combined analyses which comprise 3307 characters (419 characters for ITS, 824 characters for LSU, 990 characters for SSU, 1074 characters for TEF1-a) after alignment. Staurosphaeria rhamnicola (MFLUCC 17-0813, MFLUCC 17-0814) (Montagnulaceae, Pleosporales) are used as the outgroup taxa. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of - 29921.960589 is presented. The matrix had 1190 distinct alignment patterns, with 23.56% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.243950, C = 0.233656, G = 0.265190, T = 0.25720; substitution rates AC = 1.334366, AG = 3.046037, AT = 2.327147, CG = 0.800430, CT = 6.652827, GT = 1.000000; gamma distribution shape parameter a = 0.527656. Maximum parsimony analysis resulted 3276 constant characters and 629 informative characters. Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 65% are placed above and below the branches respectively. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.90 are in bold. The extype strains are in bold and black. The newly generated sequences are indicated in bold and blue 123 Fungal Diversity (2019) 96:1–242 61 123 62 related to the family Hermatomytaceae with relatively strong statistical support. Pseudoberkleasmium Tibpromma & K.D. Hyde Notes: The monotypic genus Pseudoberkleasmium was introduced by Tibpromma et al. (2018) with P. pandanicola Tibpromma & K.D. Hyde as the type species. Pseudoberkleasmium is characterized by hyaline subglobose conidiogenous cells and acrogenous, broadly ellipsoidal to obovoid, muriform, brown or olivaceous green, and guttulate conidia. In this paper, we introduce a new Pseudoberkleasmium species, P. chiangmaiense, based on phylogenetic and morphological evidence. Pseudoberkleasmium chiangmaiense Y.Z. Lu & K.D. Hyde, sp. nov. Index Fungorum number: IF555595; Facesoffungi number: FoF05310; Fig. 39 Etymology: chiangmaiense, referring to collecting site. Holotype: MFLU 17-1118 Saprobic on decaying wood. Sexual morph Undetermined Asexual morph Hyphomycetous, dictyosporous. Colonies on natural substratum sporodochia, superficial, compact, scattered, irregular, black, glistening. Mycelium immersed in the substrate, composed of septate, branched, smooth, thin-walled, hyaline to pale brown, 2 lm wide hyphae. Conidiophores micronematous, mononematous, reduced, hyaline, smooth-walled. Conidiogenous cells 12– 18 9 12–18 lm ( x = 15 9 14 lm, n = 20), holoblastic, monoblastic, integrated, terminal, determinate, subglobose, with guttulate, hyaline. Conidia 30–35 9 15–20 lm Fig. 38 continued 123 Fungal Diversity (2019) 96:1–242 ( x = 33 9 18 lm, n = 20), acrogenous, solitary, broadly ellipsoidal to obovoid, flattened, muriform, guttulate, smooth-walled, brown, usually with conidiogenous cell attached. Culture characteristics: Conidia germinating on water agar and germ tubes produced from conidia within 12 h. Colonies growing on PDA, circular, with flat surface, edge entire, reaching 28 mm in 3 weeks at 28 C, white to pale brown in PDA medium. Mycelium superficial and partially immersed, branched, septate, hyaline to pale brown, smooth. Material examined: THAILAND, Chiang Mai Province, Mae Taeng District, Mushroom Research Center, on decaying wood, 25 January 2017, Yong-Zhong Lu, MRC 23 (MFLU 17-1118, holotype), ex-type living culture (MFLUCC 17-1809). GenBank numbers: ITS: MK131259, LSU: MK131260, TEF1-a: MK131261. Notes: Pseudoberkleasmium chiangmaiense is similar to P. pandanicola in having hyaline subglobose conidiogenous cells and acrogenous, broadly ellipsoidal to obovoid, muriform, brown and guttulate conidia, but differs by its larger conidiogenous cells (12–18 9 12–18 lm vs 5–11 9 9–12 lm). Phylogenetically, Pseudoberkleasmium chiangmaiense shares a sister relationship to P. pandanicola with good bootstrap support (100% ML/1.00 BYPP) (Fig. 40); the phylogeny also indicates that they are distinct species. Fungal Diversity (2019) 96:1–242 63 Fig. 39 Pseudoberkleasmium chiangmaiense (MFLU 17-1118, holotype). a Colonies on substrate. b–j Conidia. k Germinating conidium. l, m Colony on PDA (l from above view, m from below view). Scale bars: a = 200 lm, b–k = 20 lm, l, m = 20 mm Pyrenochaetopsidaceae Valenzuela-Lopez, Crous, Cano, Guarro & Stchigel Notes: ‘Cucurbitariaceae’ was revealed to represent five distinct families, i.e. the newly erected Neopyrenochaetaceae, Parapyrenochaetaceae, Pseudopyrenochaetaceae and Pyrenochaetopsidaceae, and Cucurbitariaceae sensu stricto with a narrowed concept (Valenzuela-Lopez et al. 2018). Pyrenochaetopsis is the 123 64 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 40 Phylogram generated from maximum likelihood analysis based on combined LSU, SSU, ITS, TEF1-a, and RPB2 sequence data representing Pseudoberkleasmiaceae and the closely related families in Pleosporales order. Related sequences are taken from Tibpromma et al. (2018). Sixty-four strains are included in the combined analyses which comprise 4340 characters (842 characters for LSU, 1030 characters for SSU, 512 characters for ITS, 926 characters for TEF1-a, 1030 characters for RPB2) after alignment. Lophiostoma crenatum (CBS 629.86) and L. arundinis (CBS 621.86) in Lophiostomataceae (Pleosporales) are used as the outgroup taxa. Single gene analyses were also performed to compare the topology and clade stability with combined gene analyses. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood values of - 38637.764110 is presented. The matrix had 1968 distinct alignment patterns, with 33.91% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.245895, C = 0.251442, G = 0.271420, T = 0.231244; substitution rates AC = 1.592217, AG = 3.496391, AT = 1.418709, CG = 1.209916, CT = 8.405355, GT = 1.000000; gamma distribution shape parameter a = 0.569582. Bootstrap values for maximum likelihood (ML) equal to or greater than 70% and clade credibility values greater than 0.90 (the rounding of values to 2 decimal proportions) from Bayesian-inference analysis labelled on the nodes. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue type genus of Pyrenochaetopsidaceae. The general characters of the family Pyrenochaetopsidaceae included glabrous or setose pycnidia with apapillate or papillate ostiolate necks, phialidic conidiogenous cells and aseptate, cylindrical to allantoid conidia. Members of Pyrenochaetopsidaceae are saprobes from various niches, such as water, plants, soil and air samples (de Gruyter et al. 2010, 2013; Valenzuela-Lopez et al. 2018). Some species e.g. Pyrenochaetopsis poae, have been repeatedly isolated as endophytes of graminaceous plants (Crous et al. 2014). Other species such as P. globosa, P. paucisetosa and Neopyrenochaetopsis hominis, have been isolated from sinusitis, toe nail and skin tissue of humans as weak pathogens (Valenzuela-Lopez et al. 2018). Pyrenochaetopsis Gruyter, Aveskamp & Verkley Notes: Pyrenochaetopsis was erected by de Gruyter et al. (2010) with P. leptospora (Sacc. & Briard) Gruyter, Aveskamp & Verkley as the type species (de Gruyter et al. 2010). The genus is characterized by setose pycnidia with apapillate or papillate ostioles, and simple, aseptate, cylindrical to allantoid conidia. Pyrenochaetopsis sinensis G.S. Li, J.M. Liang & L. Cai, sp. nov. Index Fungorum number: IF556011; Facesoffungi number: FoF05965; Fig. 41 Etymology: Named after the country where the type was collected. Holotype: HMAS 248045 65 Saprobic on rhizosphere soil of turfgrasses. Sexual morph Undetermined. Asexual morph Hyphae pale white, smooth- and thin-walled, septate, 2–5 lm wide. Conidiomata pycnidial, pale brown, solitary or confluent, immersed (OA), subglobose or globose, 100–180 9 100– 155 lm, with apapillate ostiolar neck. Pycnosclerotia forming after 25 days culturing on OA, brown to dark brown, globose or subglobose, 55–60 lm in diam. Setae absent. Pycnidial wall of textura angularis, 3–5 layered, 15–25 lm thick, composed of brown, flattened polygonal cells of 3–6 lm diam. Conidiogenous cells phialidic, hyaline, smooth-walled, subglobose, 3–6 9 2–5 ( x = 4 9 3, n = 30) lm. Conidia aseptate, hyaline, smooth- and thin-walled, cylindrical to allantoid, 3– 4.5 9 1–2 ( x = 4 9 2, n = 40) lm, usually with two guttules. Culture characteristics: Colonies on OA reaching 23 mm diam. after 7 d at 25 ± 1 C, flattened, greenish grey (26B2); reverse olive grey (26E3). Colonies on MEA reaching 19 mm diam. after 7 d at 25 ± 1 C, floccose, pale grey (26B1) to olive (26C2); reverse olive (26F3). NaOH spot test negative. Crystals absent. Material examined: CHINA, Beijing, in rhizosphere soil of Poa pratensis (Poaceae), 26 August 2017, J.M. Liang (HMAS 248045, holotype), ex-type living culture CGMCC 3.19296 = LC12199; Paratypes: China. Beijing, in rhizosphere soil of Digitaria sanguinalis (Poaceae), 21 August 2017, J.M. Liang LC12200); China, Beijing, in rhizosphere soil of Poa pratensis (Poaceae), 21 August 2017, J.M. Liang LC12197; ibid. LC12198. GenBank numbers: ITS: MK348586, MK348587, MK348584, MK348584, LSU: MK348581, MK348580, MK348582, MK348583, RPB2: MK355077, MK355078, TUB2: MK348221, MK348220, MK348222, MK348223. Notes: Pyrenochaetopsis sinensis is phylogenetically distinct from other species in this genus, clustering sister to P. microspore (Fig. 42). Pyrenochaetopsis sinensis is easily distinguished from P. microspora by the absence of setae covering on the surface of pycidium (Gruyter and Boerema 2002). In addition, P. sinensis is the only species of Pyrenochaetopsis producing pycnosclerotia. Tetraplosphaeriaceae Kaz. Tanaka & K. Hiray. Notes: The family Tetraplosphaeriaceae accommodates seven genera viz. Ernakulamia Subram. (Subramanian 1994), Polyplosphaeria Kaz. Tanaka & K. Hiray. (Tanaka et al. 2009), Pseudotetraploa Kaz. Tanaka & K. Hiray. (Tanaka et al. 2009), Quadricrura Kaz. Tanaka, K. Hiray. & Sat. Hatak. (Tanaka et al. 2009), Shrungabeeja V.G. Rao & K.A. Reddy (Rao and Reddy 1981), Tetraploa Berk. & Broome, and Triplosphaeria Kaz. Tanaka & K. Hiray. (Tanaka et al. 2009). Ernakulamia is placed in this family based on morphological and phylogenetic analyses 123 66 Fungal Diversity (2019) 96:1–242 Fig. 41 Pyrenochaetopsis sinensis (HMAS 248045, holotype). a– c Colony on OA, MEA and SNA (front and reverse). d Pycnidia forming on OA. e Pre-pycnidial structures. f Pycnidia. g, h Pycnidia wall. i1–i5 Conidiogenous cell. j Conidia. k Pycnosclerotium on OA. l–m Pycnosclerotium. Scale bars: e = 50 lm, g = 20 lm (apply to l), h = 10 lm (apply to i, j, m, i1–i5) (Delgado et al. 2017). Although Pseudotetraploa and Shrungabeeja were not included in this family by Wijayawardene et al. (2018a), these genera should be retained. Pseudotetraploa is distinct from Tetraploa by its obpyriform to long obpyriform, pseudoseptate conidia composed of 4 to 8 columns with 4 (rarely 6 to 8) setose appendages (Hyde et al. 2013). Pseudotetraploa also formed an independent clade in the molecular analysis and distinguished from Tetraploa (Tanaka et al. 2009). Shrungabeeja was identified as a distinctive genus within Tetraplosphaeriaceae in Ariyawansa et al. (2015). Tetraploa, Polyplosphaeria and Triplosphaeria are characterized with almost hyaline 1(–3)-septate ascospores as Massarina-like sexual morphs and several setose 123 Fungal Diversity (2019) 96:1–242 67 Fig. 42 Phylogram generated from maximum likelihood (ML) analysis inferred from a four-locus concatenated alignment (ITS, LSU, RPB2 and TUB2) for Pyrenochaetopsis species and closely related species. Related sequences are taken from Valenzuela-Lopez et al. (2018) and Crous et al. (2014). Twenty-two strains are included in the multi-locus analyses which comprise a total of 2454 characters (443 characters for ITS, 817 characters for LSU, 297 characters for TUB2 and 897 characters for RPB2) after alignment. Neopyrenochaetopsis hominis (CBS 143033) (Pyrenochaetopsidaceae, Pleosporales) is used as the outgroup taxon. Tree topology of the ML analysis is similar to that of the Bayesian analysis. The best RAxML tree with a final likelihood value of - 8929.890374 is presented. The matrix had 573 variable sites. Estimated base frequencies were as follows: A = 0.2382, C = 0.2531, G = 0.2663, T = 0.2423; substitution rates AC = 1.4907, AG = 4.6367, AT = 1.4907, CG = 1.0000, CT = 10.5414, GT = 1.0000; gamma distribution shape parameter a = 0.5740. Bootstrap values (1000 replicates) over 75% for ML and Bayesian posterior probability (PP) over 0.95 are marked at the left of every node (ML/BYPP). Type strains are in bold and the newly generated sequences are indicated in blue appendages as tetraploa-like asexual morphs (Tanaka et al. 2009; Hyde et al. 2013). The sexual morphs of the genera Ernakulamia, Pseudotetraploa, Shrungabeeja and Quadricrura are unknown. Most taxa of Tetraplosphaeriaceae were found on bamboo and grasses. are 20 epithets listed under the genus in Index Fungorum (2019). However, only five species of Tetraploa have molecular data (NCBI: www.ncbi.nlm.nih.gov; accessed 1st May 2019). Tetraploa has relatively small globose ascomata, narrowly fusiform ascospores having appendage-like sheath, and conidia with four setose appendages and is mainly known from monocotyledons especially grasses and bamboo (Tanaka et al. 2009; Goonasekara et al. 2018; Farr and Rossman 2019). A new record of Tetraploa nagasakiensis from China is reported in this paper. Tetraploa Berk. & Broome Notes: Hyde et al. (2013) and Wijayawardene et al. (2014) recommended the use of Tetraploa over Tetraplosphaeria based on the close phylogenetic relationships between the type species of these two genera and there are more species accommodated in Tetraploa, having priority of the most named uses. Thus, Rossman et al. (2015b) adopted Tetraploa over Tetraplosphaeria. There Tetraploa nagasakiensis (Kaz. Tanaka & K. Hirayama) Kaz. Tanaka & K. Hirayama 123 68 Fungal Diversity (2019) 96:1–242 Facesoffungi number: FoF05080; Fig. 43 Basionym: Tetraplosphaeria nagasakiensis Kaz. Tanaka & K. Hiray., in Tanaka et al., Stud. Mycol. 64: 180 (2009) Saprobic on dead bamboo branches. Sexual morph Ascomata 170–200 lm high, 190–260 lm diam., immersed under host epidermis, solitary, scattered, visible as small minute black spots or papilla on host issue, dark Fig. 43 Tetraploa nagasakiensis (HKAS 101756, new geographical record). a–c Host and ascomata. d Vertical section of ascoma. e Peridium of ascoma. f Pseudoparaphyses. g–k Asci. l Culture from above and below views. m, n Ascospores. o Germinating ascospore. Scale bars: i–k, o = 20 lm, e, g, h = 15 lm, f, m, n = 10 lm 123 Fungal Diversity (2019) 96:1–242 brown to black, uni-loculate, globose to subglobose or ampulliform, coriaceous, with ostiole. Peridium 17–24 lm wide, comprising several layers, with outers layers composed of dark brown to black cells of textura angularis, inner layers composed of light brown to hyaline pseudoparenchymatous cells arranged in textura angularis. Hamathecium of 1.7–2.4 lm, septate, branched, broad pseudoparaphyses, anastomosing near the apex. Asci 98– 130 9 14–18 lm ( x = 114 9 16 lm, n = 20), 8-spored, bitunicate, fissitunicate, clavate, pedicellate with furcate ends, apically rounded with well-developed ocular chamber. Ascospores 30–34.5 9 4–5 lm ( x = 32.3 9 4.6 lm, n = 20), overlapping 2–3-seriate, hyaline, 1-septate, narrowly fusiform, smooth-walled, multi-guttulate. Asexual morph Refer to Tanaka et al. (2009). Culture characteristics: Ascospores germinating on PDA within 24 h. Colonies on PDA reaching 30–35 mm diam. after 4 weeks at 20–25 C, circular, slightly raised to umbonate, pale black at the middle, white at the margin from the above and below, no pigment in agar. Material examined: CHINA, Yunnan Province, Honghe Prefecture, Pingbian County, nearby Tuanpo Reservoir, on dead bamboo branches, 21 September 2017, H.B. Jiang, Pb001 (HKAS 101756), living culture (KUMCC 18-0109). GenBank numbers: ITS: MK079890, LSU: MK079891, SSU: MK079888. Notes: Although the new collection (HKAS 101756) has eight different ITS base pairs as compared to the type strain of Tetraploa nagasakiensis, they are still close to each other in multi-phylogenetic analysis based on ITS, LSU and SSU sequenced data (Fig. 44). They are also morphologically very similar. Tetraploa nagasakiensis was reported only from Japan (Tanaka et al. 2009), but our new strain was collected in China. Torulaceae Corda Notes: The family Torulaceae was introduced by Corda (Sturm 1829). Two genera, Dendryphion Wallr. and Torula Pers., are accepted within this family. Su et al. (2016) and Li et al. (2016) introduced the genera Neotorula Ariyaw., Z.L. Luo & K.D. Hyde and Sporidesmioides Jin F. Li, Phook. & K.D. Hyde within Torulaceae, respectively. Torula Pers. Notes: The genus Torula was established by Persoon (1794), typified by T. herbarum (Pers.) Link as the type species. Torula is characterized by terminal or lateral, monoblastic or polyblastic conidiogenous cells, which have a basally thickened and heavily melanized wall, with the apex thin-walled and frequently collapsing and becoming coronate; these conidiogenous cells are consequently termed ‘corona cells’ (Crane and Miller 2016). This genus is problematic as many species await redescription (Seifert et al. 2011). More sequence data are required in the future 69 study of the genus. A key to Torula and similar genera was provided by Crane and Miller (2016). Torula breviconidiophora C.G. Lin & K.D. Hyde, sp. nov. Index Fungorum number: IF555428; Facesoffungi number: FoF05069; Fig. 45 Etymology: In reference to the short conidiophores. Holotype: MFLU 18-1693 Saprobic on submerged decaying wood. Sexual morph Undetermined. Asexual morph Colonies effuse on the natural substrate, scattered, hairy, dark brown. Mycelium immersed to superficial, composed of pale brown, septate, branched hyphae. Conidiophores macronematous, mononematous, subcylindrical, erect, septate, smooth, straight or flexuous, subhyaline to dark brown, with ampulliform cells, 3.5–28 lm long, 3–8 lm wide at the widest part. Conidiogenous cells mono- to polyblastic, integrated, terminal, on conidiophores, terminal or intercalary in conidial chains, doliiform to ellipsoid, pale brown, collapsing in above half when old, 4–7 lm long, 3–7 lm wide. Conidia phragmosporous, in branched chains, acrogenous, dry, cylindrical, medium to dark brown, with apical cell subhyaline or pale brown, 1–4septate, deeply constricted at septa, verrucose, rounded at both ends, easily separating, 8–21 lm long (x ¯ = 15.2 lm, n = 60), 3.5–7 lm wide (x ¯ = 5.2 lm, n = 60). Culture characteristics: Conidia germinating on PDA within 12 h. Colonies on PDA effuse, greyish brown to dark brown, reaching a diam. of 4-6 cm in 10 days at 25 C. Material examined: THAILAND, Chang Rai Province, Muang District, Mae Fah Luang University, near S7 building, on decaying wood, 20 July 2016, Chuangen Lin, ML 1-4 (MFLU 18-1693, holotype; HKAS 102196, isotype), ex-type living culture (KUMCC 18-0130). GenBank numbers: ITS: MK071670, LSU: MK071672, SSU: MK071697, TEF1-a: MK077673. Notes: This species is most similar to T. mackenziei which is characterized by greyish brown conidia composed of moniliform cells and 2–3 septa (Li et al. 2017). However, the conidia of T. breviconidiophora are medium to dark brown, with 1–4-septate, subhyaline or pale brown apical cells. From the phylogenetic trees of Bayesian inference, maximum parsimony and maximum likelihood analyses based on combined ITS, LSU, SSU and TEF1-a, this species is closest to T. chromolaenae and T. mackenziei (Fig. 47). Torula chromolaenae is distinct from other Torula species in having smaller and doliiform conidiogenous cells and 2–3-septate conidia (Li et al. 2017). Torula breviconidiophora differs from T. chromolaenae by its longer conidiophores (3.5–28 lm vs 5–6.3 lm). 123 70 Fig. 44 Phylogram generated from maximum likelihood analysis based on a combined ITS, LSU, and SSU sequence dataset for Tetraploa species and several closely related genera in Tetraplosphaeriaceae. Related sequences are taken from Tanaka et al. (2009) and Ariyawansa et al. (2015). Twenty-nine strains are included in the combined genes sequence analyses which comprise total 3225 characters including gaps (584 characters for ITS, 1300 characters for LSU, 1341 characters for SSU) after alignment. Lophiotrema neoarundinaria KT 856 (Lophiotremataceae, Pleosporales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. 123 Fungal Diversity (2019) 96:1–242 The best sorting RaxML tree with a final likelihood value of - 8716.750658 is presented. The matrix had 430 distinct alignment patterns, with 15.62% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.251593, C = 0.234741, G = 0.277315, T = 0.236350; substitution rates AC = 4.295630, AG = 4.189995, AT = 2.538258, CG = 1.455605, CT = 14.145092, GT = 1.000000; gamma distribution shape parameter a = 0.501945. Bootstrap values for maximum likelihood (ML) equal to or greater than 70% are shown on the left. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.90 are shown on the right. The ex-type strains are in bold and black. The newly generated sequence is indicated in blue Fungal Diversity (2019) 96:1–242 71 Fig. 45 Torula breviconidiophora (MFLU 18-1693, holotype). a, b Host material. b Conidiophores on leaf surface. c, d Conidiophores, conidiogenous cells and conidia. e–g Conidia. Scale bars: c–g = 10 lm Torula polyseptata C.G. Lin & K.D. Hyde, sp. nov. Index Fungorum number: IF555429, Facesoffungi number: FoF05070; Fig. 46 Etymology: In reference to the 2–8-septate conidia. Holotype: MFLU 18-1694 Saprobic on submerged decaying wood. Sexual morph Undetermined. Asexual morph Colonies effuse on the natural substrate, scattered, hairy, dark brown. Mycelium immersed to superficial, composed of pale brown, septate, branched hyphae. Conidiophores macronematous, mononematous, subcylindrical, erect, septate, smooth, straight or slightly flexuous, dark brown to black, 10– 40 lm long, 3.5–8 lm wide at the widest part. Conidiogenous cells holoblastic, mono- to polyblastic, integrated, terminal or intercalary in conidial chains, doliiform, pale brown, collapsing in above half when old, 4.5–8.5 lm long, 4.5–8 lm wide. Conidia phragmosporous, in branched chains, acrogenous, dry, cylindrical, medium to dark brown, 2–8-septate, deeply constricted at septa, verrucose, rounded at both ends, easily separating, 10–40 lm long ( x = 19.3 lm, n = 35), 3.5–7.5 lm wide ( x = 5.5 lm, n = 35). Culture characteristics: Conidia germinating on PDA within 12 h. Colonies on PDA effuse, greyish brown to dark brown, reaching a diam. of 4–6 cm in 10 days at 25 C. Material examined: THAILAND, Chang Rai Province, Muang District, Mae Fah Luang University, on decaying wood, 20 July 2016, Chuangen Lin, ML 21-2 (MFLU 18-1694, holotype; HKAS 102197, isotype), ex-type living culture (KUMCC 18-0131). GenBank numbers: ITS: MK071671, LSU: MK071673, SSU: MK071698, TEF1-a: MK077674. Notes: This species is most similar to Torula pluriseptata in their conidial morphology, however, the conidiophores of T. polyseptata are longer than T. pluriseptata 123 72 Fungal Diversity (2019) 96:1–242 Fig. 46 Torula polyseptata (MFLU 18-1694, holotype). a, b Host material. b Conidiophores on leaf surface. c–e Conidiophores, conidiogenous cells and conidia. f–j Conidia. Scale bars: c = 20 lm, d–j = 10 lm (10–40 lm vs 2.8–4.3 lm). Torula polyseptata forms a clade together with strains of T. chiangmaiensis with 98% MP bootstrap support, 97% ML bootstrap support and 0.97% Bayesian posterior probabilities within the genus Torula (Fig. 47). Torula polyseptata is distinct from T. chiangmaiensis by the size and number of septa in the conidia. The ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue. The tree is rooted with Dendryphion nanum (HKAS 84010). Trematosphaeriaceae K.D. Hyde, Y. Zhang ter, Suetrong & E.B.G. Jones Notes: Trematosphaeriaceae was introduced by Suetrong et al. (2011b) to accommodate the genera 123 Falciformispora K.D. Hyde, Halomassarina Suetrong et al. and Trematosphaeria Fuckel. These genera mainly comprise fungi inhabiting mangrove wood in marine environments and wood in freshwater (Hyde 1989; Suetrong et al. 2011b). The main distinguishing characters of this family are medium-sized rounded ascomata with a papillate ostiole, a relatively wide, coriaceous peridium, cellular pseudoparaphyses and cylindro-clavate asci. The ascospores are two-celled or many celled, hyaline or brown. We introduce Falciformispora aquatica as a new species based on evidence from morphology and phylogenetic analyses of combined LSU, SSU, RPB2 and TEF1-a sequence data (Fig. 49). Fungal Diversity (2019) 96:1–242 73 Fig. 47 Phylogenetic tree generated from ML analysis based on combined ITS, LSU, SSU and TEF1-a sequence data for the genus Torula. Related sequences are taken from Crane and Miller (2016), Su et al. (2016) and Li et al. (2017). Twenty-eight strains are included in the combined analyses which comprise 3139 characters (552 characters for ITS, 844 characters for LSU, 887 characters for SSU, 856 characters for TEF1-a) after alignment. The best RaxML tree with a final likelihood value of - 7637.920840 is presented. The matrix had 454 distinct alignment patterns, with 31.43% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.242476, C = 0.262534, G = 0.274745, T = 0.220245; substitution rates AC = 3.217101, AG = 3.132488, AT = 2.857294, CG = 1.885404, CT = 12.537654, GT = 1.000000; gamma distribution shape parameter a = 0.02. Maximum parsimony analysis of 2817 constant characters and 204 informative characters resulted in 126 equally most parsimonious tree of 567 steps (CI = 0.704, RI = 0.785, RC = 0.552, HI = 0.296). For the Bayesian analysis, two parallel runs with six chains were run for 1000,000 generations and trees were sampled every 100th generation, resulted in 20002 trees from two runs of which 15002 trees were used to calculate the posterior probabilities (each run resulted in 10001 trees of which 7501 trees were sampled). Bootstrap support values for maximum parsimony (MP) and maximum likelihood (ML) greater than 50% and Bayesian posterior probabilities greater than 0.8 are indicated above or below the nodes as MLBS/MPBS/BYPP Falciformispora K.D. Hyde Notes: Falciformispora was established by Hyde (1992) as a monotypic genus and was assigned to the Pleosporaceae when compared with Setosphaeria, which has an Exserohilum asexual morph and is exclusively parasitic on Gramineae. Subsequently, Raja et al. (2008) collected Falciformispora species in freshwater in Florida and found that it is closely related to Chaetomastia rather than Setosphaeria, but is distinct in having hyaline ascospores. Suetrong et al. (2009) recorded it from the terrestrial oil palm (Elaeis guineensis) in Thailand. Suetrong et al. (2011b) showed that Falciformispora forms a well-supported clade with Trematosphaeria pertusa and Halomassarina thalassiae in Trematosphaeriaceae based on phylogenetic analysis. Therefore, this genus was placed in Trematosphaeriaceae. Currently, F. lignatilis, F. senegalensis and F. tompkinsii are listed in Index Fungorum (2019). Falciformispora aquatica D.F. Bao, K.D. Hyde & H.Y. Su, sp. nov. Index Fungorum number: IF555415; Facesoffungi number: FoF05059; Fig. 48 123 74 Fungal Diversity (2019) 96:1–242 Fig. 48 Falciformispora aquatica (MFLU 18-1228, holotype). a Ascomata on submerged wood. b Section of ascoma. c, d Section of peridium. e Pseudoparaphyses. f–k Asci. l-n Ascospores. o Germinating ascospore. p, q Culture on PDA. Scale bars: b = 100 lm, f–k = 30 lm, c–e, l–o = 20 lm Etymology: Referring to the aquatic habitat of this fungus. Holotype: MFLU 18-1228 Saprobic on decaying wood, submerged wood in freshwater. Asexual morph Undetermined. Sexual morph Ascomata 260–330 lm diam. 230–300 lm high ( x = 286 9 260 lm, n = 10) solitary to gregarious, immersed in the host tissue, globose to subglobose, coriaceous, papillate, ostiolate, dark grey to black. Peridium up to 25–30 lm wide, composed of outer thick-walled angular or rounded greyish brown cells, and inner hyaline cells of textura prismatica. Hamathecium comprising 2– 3.5 lm wide, numerous, septate, cellular pseudoparaphyses, hyaline, branched, surrounded by a gelatinous matrix. Asci 115–167 9 28–47 lm ( x = 147 9 36 lm, n = 20), 8spored, bitunicate, fissitunicate, cylindro-clavate, short 123 Fungal Diversity (2019) 96:1–242 pedicellate. Ascospores 48–59 9 10–16 lm ( x = 55 9 13 lm, n = 30), fusiform to clavate, hyaline, straight or slightly curved, 5–6-septate, mostly 6-septate, slightly constricted at all septa, tapering to narrow both of ends, surrounded by a mucilaginous sheath and a single scythe-like appendage at the apex. Culture characteristics: Ascospores germinating on PDA within 24 h. Colonies on MEA at room temperature reaching 2.5 cm diam. in three weeks, mycelium pale brown to grayish brown after 3 weeks, composed brown to dark brown, septate, smooth or verrucose hyphae. Material examined: Thailand, Prachuap Khiri Khan Province, Sai Khu Waterfall, on submerged decaying wood, August 2017, Vinit Kumar., site1–2–1, (MFLU 18–1228, holotype), ex-type living culture (MFLUCC 18–0212). GenBank numbers: LSU: MK063643, RPB2: MK099810, TEF1-a: MK099811. Notes: The phylogenetic analyses showed that Falciformispora aquatica grouped with members of Falciformispora, but separated in an independent lineage with relatively high bootstrap support (100% ML/1.00 BYPP, Fig. 49). Falciformispora aquatica is similar to F. lignatilis, as they have similar asci (bitunicate, fissitunicate, with a short pedicel) and ascospores (fusiform to clavate, hyaline). However, F. aquatica has larger asci (115–167 9 28–47 vs. 110–136 9 20–32 lm), and larger ascospores (48–59 9 10–16 vs. 42–50 9 7.5–10 lm). Moreover, F. aquatica ascospores are mostly 6-septate, whereas those of F. lignatilis over 7-septate. Minutisphaerales Raja, Oberlies, Shearer & A.N. Mill. Notes: The order Minutisphaerales was introduced by Raja et al. (2015) based on a LSU, ITS, SSU and MCM7 sequence dataset. It is typified by the sexual genus Minutisphaera with the chemical analysis carried out for M. aspera Raja, Oberlies, Shearer & A.N. Mill. and M. parafimbriatispora Raja, Oberlies, Shearer & A.N. Mill. Jayasiri et al. (2018) included a new family Acrogenosporaceae in Minutisphaerales. Acrogenosporaceae Jayasiri & K.D. Hyde Notes: Acrogenosporaceae is a monotypic family with Acrogenospora as the type genus. Rossman et al. (2015b) suggested to protect Acrogenospora over Farlowiella (Art. 59.1). Ellis (1972) reported an Acrogenospora asexual morph accompanies F. australis Dennis on the host substrate, while Mason (1941) showed the connection between A. megalospora (Berk. & Broome) Goh, K.D. Hyde & C.K.M. Tsui and F. armichaeliana (Berk.) Sacc. based on a cultural study. With seven strains related to this group, Jayasiri et al. (2018) carried out the phylogenetic analyses which supported the connection between Acrogenospora and Farlowiella. 75 Acrogenospora M.B. Ellis Notes: The holomorph genus Acrogenospora is characterized by laterally compressed hysterothecia with a prominent sunken slit, aseptate, hyaline pedicellate ascospores, macronematous brown conidiophores, monoblastic, terminal or intercalary conidiogenous cells and globose, elliposoid or obovoid, olivaceous to brown conidia. Among the available sequenced strains of the genus, many do not have any associated morphological descriptions. We therefore designate a reference specimen for the type species A. sphaerocephala and introduce a new taxon, A. thailandica, based on phylogenetic analyses and morphological characters. Acrogenospora sphaerocephala (Berk. & Broome) M.B. Ellis, Dematiaceous Hyphomycetes: 114 (1971) : Monotospora sphaerocephala Berk. & Broome, Ann. Mag. nat. Hist., Ser. 3 3: 361 (1859) Index Fungorum number: IF308236; Facesoffungi number: FoF04687; Fig. 50 Saprobic on submerged decaying wood. Sexual morph Undetermined. Asexual morph Colonies sparse, scattered, black, glistening, hairy. Mycelium mostly immersed, consisting of septate, thin-walled, smooth, hyaline to pale brown hyphae. Conidiophores macronematous, mononematous, (155–)215–320(–360) 9 (3.5–)4.8–7.5(–9.5) lm ( x = 275 9 6 lm, n = 15), solitary, erect, mostly flexuous, septate, slightly tapering towards the apex, mid to dark brown, paler and rounded at the apex, smooth, thick-walled at the base. Conidiogenous cells monoblastic, integrated, intercalary, mid brown, cylindrical, with percurrent proliferations, sometimes flexuous at the proliferation. Conidia acrogenous, 18–30 lm ( x = 24 lm, n = 40) diam., holoblastic, olive-green to brown, spherical or subspherical, unicellular, smooth and thick-walled, guttulate, truncate at the base. Culture characteristics: Conidia germinating on PDA within 24 h. Germ tubes produced from both ends. Colonies on PDA, reaching 5–10 mm diameter after two weeks at 25 C in natural light, with dense grayish green aerial mycelium on the surface, black in reverse with entire margin. Material examined: THAILAND, Prachuap Khiri Khan Province, near 1230.1950 N, 9931.3500 E, on decaying wood submerged in a freshwater stream, 25 December 2014, Jaap van Strien, Site 5-14-2 (MFLU 18-1130, reference specimen designated here), living culture (MFLUCC 16-0179, GZCC 15-0071). GenBank numbers: ITS: MH606233, LSU: MH606222, RPB2: MH626448. Notes: The genus Acrogenospora includes 11 species and two of them have been linked to sexual morphs in Farlowiella. The synonymy of Acrogenospora with 123 76 Fungal Diversity (2019) 96:1–242 Fig. 49 Phylogram generated from maximum likelihood analysis based on combined LSU, SSU, TEF1-a and RPB2 sequence data for Trematosphaeriaceae and several closely related families in Dothideomycetes. Related sequences are taken from Suetrong et al. (2009) and Wanasinghe et al. (2018a), forty-three strains are included in the combined analyses which comprise total 3628 characters (1119 characters for LSU, 983 characters for SSU, 821 characters for TEF1a, 705 characters for RPB2) after alignment. Pleospora herbarum (CBS 191.86) (Pleosporaceae, Pleosporales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of –24083.791079 is presented. The matrix had 1258 distinct alignment patterns, with 18.62% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.250098, C = 0.240190, G = 0.272034, T = 0.237678; substitution rates AC = 1.254339, AG = 3.593578, AT = 1.238765, CG = 1.071163, CT = 7.108157, GT = 1.000000; gamma distribution shape parameter a = 0.194764. Bootstrap values for maximum likelihood (ML) and equal to or greater than 75% and Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are placed above the branches. The newly generated sequence is indicated in bold and blue Farlowiella has long been accepted based on the distinctive morphology of the asexual morphs (Ellis 1971, 1976; Goh et al. 1998a; Rossman et al. 2015b). The generic name Acrogenospora has been protected over its sexual morph Farlowiella (Rossman et al. 2015b). In this study, molecular data has been generated to support its asexual-sexual connection. Acrogenospora sphaerocephala was designated as the type species in the genus (Ellis 1971) and it is a frequently observed species with a worldwide distribution (Goh et al. 1998a). Phylogenetic analyses indicate that our collection A. sphaerocephala (MFLUCC 16-0179) formed a distinctive branch in the genus, close to A. carmichaeliana (FMR 11021, CBS 179.73, CBS 164.76 and CBS 206.36). However, the identification of these A. carmichaeliana (Berk.) Rossman & Crous collections is questionable as only molecular sequence data was provided and sequence data is unavailable for the ex-type strain. Acrogenospora sphaerocephala was thought to be the asexual morph of F. carmichaeliana due to the information in GenBank. It must be wrongly identified as A. megalospora was introduced as the asexual morph of Farlowiella carmichaeliana (Ellis 1971; Goh et al. 1998a) and the combination of the holomorph species was suggested as Acrogenospora carmichaeliana (Rossman et al. 2015b). Our collection was identified as A. sphaerocephala with the morphological characters well-matched with the original diagnosis. Acrogenospora sphaerocephala is morphologically similar to A. megalospora except A. sphaerocephala has globose conidia, while A. megalospora produces broadly ellipsoidal 123 Fungal Diversity (2019) 96:1–242 Fig. 50 Acrogenospora sphaerocephala (MFLU 18-1130, reference specimen). a Colony on substrate. b, c Conidiophores with conidia. d, e Conidiophores. f, g Conidiogenous cells with conidia. h–j Conidia. 77 k Germinated conidium on PDA. l, m Culture (l from above view, m from below view). Scale bars: a = 200 lm, b–d = 100 lm, e– g = 50 lm, h = 30 lm, k = 20 lm, i, j = 10 lm 123 78 or obovoid conidia (Ellis 1971; Goh et al. 1998a; Cai et al. 2006). Additionally, our collection has similar size of conidiophores (155–360 9 3.5–9.5 lm) with given by Ellis (1971) (up to 380 lm long, 5–8 lm wide at the apex and 9–11 lm wide at the base) and similar conidial size (18–30 lm, x = 24 lm) with given by Ellis (1971) (15– 33 9 14–33 lm, x = 28 9 27 lm). Thus, a reference specimen for A. sphaerocephala is designated here. Acrogenospora thailandica J. Yang & K.D. Hyde, sp. nov. Index Fungorum number: IF555499; Facesoffungi number: FoF04676; Fig. 51 Etymology: Referring to the distribution in Thailand. Holotype: MFLU 18-1129 Saprobic on submerged decaying wood. Sexual morph Undetermined. Asexual morph Colonies sparse, scattered, dark brown, glistening, hairy. Mycelium mostly immersed, consisting of septate, hyaline to pale brown hyphae. Conidiophores 850–950 9 3.5–8 lm, macronematous, mononematous, solitary, erect, straight or slightly flexuous, septate, long cylindrical, slightly tapering towards the apex, pale to dark brown, paler towards the apex, smooth, thickwalled at the base. Conidiogenous cells monoblastic, integrated, intercalary, brown, cylindrical. Conidia acrogenous, 15.5–24.5 lm ( x = 19 lm, n = 35) diam., holoblastic, olive-green to dark brown, spherical or subspherical, unicellular, smooth and thick-walled, guttulate, truncate at the base. Culture characteristics: Conidia germinating on PDA within 24 h. Germ tubes produced from both ends. Colonies on PDA slow growing, reaching 7–10 mm diameter after two weeks at 25 C in natural light, circular, with dark grayish green aerial mycelium on the surface, black in reverse with entire margin. Material examined: THAILAND, Trat Province, Amphoe Ko Chang, 12080 N, 102380 E, on decaying wood submerged in a freshwater stream, 27 April 2017, YongZhong Lu, YJT-30-2 (MFLU 18-1129, holotype; HKAS 102140, isotype), ex-type living culture (MFLUCC 17-2396). GenBank numbers: ITS: MH606234, LSU: MH606223, RPB2: MH626449, SSU: MH606221. Notes: Phylogenetic analyses indicate that Acrogenospora thailandica forms a basal branch in the genus together with its sister taxon Acrogenospora sp. (JX-43) with strong support. Yang et al. (2016c) reported Acrogenospora sp. (JX-43) with LSU, SSU and ITS sequence data and only mentioned that Acrogenospora sp. (JX-43) is morphologically similar to A. sphaerocephala. We therefore identify Acrogenospora sp. (JX-43) as A. thailandica based on the phylogenetic result (Fig. 52). Acrogenospora thailandica resembles A. sphaerocephala in having relatively long conidiophores, monoblastic, intercalary conidiogenous 123 Fungal Diversity (2019) 96:1–242 cells and globose conidia of similar size, but A. thailandica has much longer conidiophores (850–950 lm long) than those of A. sphaerocephala (155–360 lm long). Dothideomycetes orders incertae sedis Asterinales M.E. Barr ex D. Hawksw. & O.E. Erikss. Notes: The order Asterinales is poorly understood due to minimal sequence data and its taxonomic placement is mainly based on morphology. There are different interpretations of this order. Hongsanan et al. (2014) included only a single family Asteraceae in Asterinales, while Guatimosim et al. (2015) added Parmulariaceae to the order. The phylogenetic placement of Asterinales are also different in the above studies. Until now, there is no convincing evidence that can support the placement of this family and it is considered as a polyphyletic group. The enrichment of sequence data of taxa of Asterinales will make contributions towards further understanding the black mildews. Asterinaceae Hansf. Notes: The family Asterinaceae was established as a member of Myriangiales by Hansford (1946). Species of Asterinaceae are characterized by dark brown hyphae with hyphopodia, dark brown thyriothecia with stellate dehiscence and dark brown, 1-septate ascospores. Phylogenetic studies have provided several different interpretations of this family (Hongsanan et al. 2014; Guatimosim et al. 2015). There is presently no convincing evidence that can support the placement of this family and it is considered as a polyphyletic group. In this study, we provide two new collections with sequences data that cluster in Asterinales sensu lato. Lembosia Lév. Notes: Lembosia was established by Leveillé (1845) with descriptions of L. dendrochili Lév., L. drimydis Lév., L. macula Lév. and L. tenella Lév. The genus was placed in Lembosiaceae by Hosagoudar et al. (2001). The genus is characterized by oval, elongate thyriothecia with X-, Y-shaped, or longitudinal dehiscence and with lateral appressoria on the hyphae. Hongsanan et al. (2014) referred the genus in Asterinaceae and treated Lembosiaceae as a synonym. Lembosia xyliae X.Y. Zeng, T.C. Wen & K.D. Hyde Facesoffungi number: FoF00933; Figs. 53, 54 Colonies epiphyllous, circular, dense, single to confluent. Hyphae superficial, straight to substraight, dark brown, branching alternate to opposite at acute to wide angles, reticulate. Hyphopodia subglobose, 1-celled, alternate, lateral, antrorse. Sexual morph Thyriothecia dense, elongate, with stellate dehiscence, 350–500 9 200–300 lm ( x = 400 9 250 lm, n = 20). Pseudoparaphyses filiform, Fungal Diversity (2019) 96:1–242 Fig. 51 Acrogenospora thailandica (MFLU 18-1129, holotype). a Colony on substrate. b, c Conidiophores with conidia. d, e Conidiogenous cells with conidia. f, g Conidia. h Germinated conidium on PDA. i Reproduced conidiophores and conidia from 79 culture. j, k Culture (j from above view, k from below view). Scale bars: a = 500 lm, b, c = 200 lm, i = 100 lm, d = 30 lm, e– h = 20 lm 123 80 Fungal Diversity (2019) 96:1–242 Fig. 52 Phylogram generated from maximum likelihood analysis based on combined LSU, SSU, TEF1-a and RPB2 sequence data for Acrogenosporaceae. Nine strains are included in the combined genes sequence analyses which comprise total 4150 characters. Hysterographium fraxini (CBS 242.32 and CBS 109.43) is selected as the outgroup taxa. Single gene analyses are carried out and the topology of each tree with clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best sorting RaxML tree with a final likelihood valued of - 8661.40414 is presented. The matrix had 269 distinct alignment patterns, with 44.96% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.259633, C = 0.231472, G = 0.273835, T = 0.235059; substitution rates AC = 27.402898, AG = 169.064131, AT = 43.635801, CG = 74.684163, CT = 684.643950, GT = 1.000000; gamma distribution shape parameter a = 0.194764. Bootstrap values for maximum likelihood (ML) and equal to or greater than 70% and Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are placed above the branches. The newly generated sequences are indicated in bold and blue dense, numerous, septate, hyaline. Asci 8-spored, bitunicate, ellipsoid to subglobose, sessile, 40–60 9 30–50 lm ( x = 50 9 40 lm, n = 30). Ascospores obovoid to ellipsoid, 2-celled, slightly constricted at the septum, lower cell slightly longer and narrower, hyaline with two oil drops in each cell when immature, becoming brown at maturity, 25– 32 9 12–17 lm ( x = 29 9 14 lm, n = 30). Asexual morph Undetermined Material examined: THAILAND, Chiang Rai Province, Doi Mae Salong, on leaves of Shorea roxburghii (Dipterocarpaceae), 22 June 2015, XY Zeng (MFLU 16-0068); THAILAND, Amphoe Ko Chang, Yuttha Navi Ko Chang Memorial, on leaves of Shorea roxburghii (Dipterocarpaceae), 27 April 2017, XY Zeng (MFLU 17-1052). GenBank numbers: ITS-LSU: MK660012 (MFLU 17-1052); LSU: MK660011 (MFLU 16-0068). Notes: This species was introduced by Ariyawansa et al. (2015) with only a single LSU sequence data. The holotype was found on leaves of Xylia sp., but it seems incorrectly identified. In this study, we sequenced both the host and the fungal species by following Zeng et al. (2018). Results indicate that the species is associated with the host Shorea roxburghii, which is very similar to Xylia species, and two new gene sequences (ITS and SSU) are generated. Lembosia shoreae (R.W. Ryan) B. Song & Hosag., which synonymized from Morenoella shoreae, is the only species reported from Shorea, but the original description is unavailable. Therefore, we would like to report a new host record and update sequence data for this species. 123 Morenoina Theiss. Notes: A recent description of Morenoina was mentioned in Tibpromma et al. (2017) where the genus was placed in Asterinales genera incertae sedis. Morenoina palmicola J. Fröhl., K.D. Hyde & Joanne E. Taylor Facesoffungi number: FoF04833; Fig. 55 Isotype: MFLU 15-0030 Saprobic on petiole of Salacca sp. Sexual morph Thyriothecia 155–895 (up to 2400) lm long, 80–190 lm wide ( x = 455 9 135 lm, n = 30), occurring on host surfaces, solitary, aggregated, or gregarious, easily removed from the host surface, superficial, ellipsoid, oblong, curved, X- or Y-shaped, flat, with longitudinal, slit-like opening, Fungal Diversity (2019) 96:1–242 81 Fig. 53 Lembosia xyliae (MFLU 16-0068, new host record). a Host leaves. b, c Thyriothecia on leaf surface. d Squash mount of thyriothecium. e Cross section of thyriothecium. f Hyphae with hyphopodia. g–i Young asci. j–m Young ascospores. Scale bars: b– d = 200 lm, e, f = 50 lm, g–m = 20 lm linear fissure, which are branched at the margin, from the centre to the outer rim, free hyphae and appressoria at the margin. Upper wall comprises linear cells, with irregular, filiform hyphae, radiating from the centre to the outer rim. Asci 19–26 9 10–12 lm ( x = 24 9 11 lm, n = 10), 8spored, bitunicate, globose to subglobose or clavate, or saccate to globose, apedicellate, with a distinct, thickened apical region. Ascospores 9–14 9 4–7 lm ( x = 12 9 6 lm, n = 20), oblong or fusiform, wider at the apex, with slightly acute ends, 1-septate, with two large guttules in each cell, hyaline, smooth-walled. Asexual morph Undetermined. Culture characteristics: Ascospore germinating on MEA within 24 h and germ tube produced from both end cells. Colonies on MEA reaching 3–4 cm diam., after two weeks, grey to olivaceous, dense, with a fairly fluffy surface, hyphae, septate, branched, and smooth-walled. Material examined: THAILAND, Krabi Province, on dead petiole of Salacca sp. (Arecaceae), 8 December 2014, Sirinapa Konta, KBR05 (MFLU 15-0030, isotype), ex-type living culture (MFLUCC 15-0284). GenBank numbers: ITS: MK120273, LSU: MK120272, SSU: MK120299. Notes: Morenoina has long taxonomic confusion concerning its familial placement. Until recently, there was no sequence data to confirm the relationships of this genus. Theissen (1913) introduced Morenoina with M. antarctica as the type species. Doidge (1942) synonymised Morenoina with Lembosia and Von Arx and Müller (1975) placed Morenoina in Leptopeltidaceae. Fröhlich and Hyde (2000) described and illustrated M. palmicola in 123 82 Fungal Diversity (2019) 96:1–242 Fig. 54 Lembosia xyliae (MFLU 17-1052, new host record). a Host leaves. b Thyriothecia on leaf surface. c Squash mount of thyriothecium. d Hyphae with hyphopodia. e Young ascus. f–i Ascospores from young to mature state. Scale bars: b, c = 100 lm, d = 50 lm, e = 20 lm, f–i = 10 lm Asterinaceae. Lumbsch and Huhndorf (2010) also suggested that its placement should be in Asterinaceae. Hongsanan et al. (2014) transferred Morenoina to Aulographaceae based on morphological characters. Tibpromma et al. (2017) introduced a new species M. calamicola and showed its unstable phylogenetic placement. There are 26 epithets of Morenoina listed in Index Fungorum (2019). We collected a fresh specimen which is similar to M. palmicola J. Fröhl. et al. but from a dead petiole of Salacca (Arecaceae) collected in Krabi, 123 Fungal Diversity (2019) 96:1–242 83 Fig. 55 Morenoina palmicola (MFLU 15-0013, new host record). a Appearance of thyrothecia on host substrate. b, c Close up of thyrothecia. d Cell walls of thyrothecium with radial arrangement. e Asci. f, g Ascospores. h, i Germinated ascospores. j, k Culture on MEA. Scale bars: a = 500 lm, b, c = 200 lm, d, e = 10 lm, f– i = 5 lm Thailand. The holotype was found on Calamus (Arecaceae) in Australia (Fröhlich and Hyde 2000). Since they have similar features it is wise to introduce our novel isolate as a new host record of Morenoina palmicola and this is the second species which has DNA based sequence data for phylogenetic analyses. In our analyses of combined LSU, and SSU sequence data, Morenoina palmicola and M. calamicola did not group in Asterinaceae sensu stricto. Both species have a close phylogenetic affiliation to Melaspileellaceae and Stictographaceae without statistical support (Fig. 56). Limited taxon sampling in the phylogenetic analyses may have resulted in inadequate resolution of this genus. Wider taxon sampling and accurate taxonomic information based on morphological examination of specimens, coupled with phylogenetic data are needed, to better integrate Morenoina into an appropriate taxonomic system. Botryosphaeriales C.L. Schoch Notes: Botryosphaeriales are considered as an important group in Dothidiomycetes due to the ecological and economic significance (Mehl et al. 2014). Many are classified as plant pathogens, known to cause many diseases on wide range of important plants. Some cause opportunistic infections in humans (de Hoog et al. 2000), and most are endophytes (Slippers & Wingfield 2007) and saprobes in plants. Based on morphological and molecular phylogenetic sequence data, Botryophaeriales comprise six families, Aplosporellaceae, Botryosphaeriaceae, Melanopsaceae, Phyllostictaceae, Planistromellaceae and Saccharataceae (Dissanayake et al. 2016, 2017a; Phillips et al. 2019). Botryosphaeriaceae Theiss. & P. Syd. Notes: Botryosphaeriaceae was introduced as a family in Botryosphaeriales by Schoch et al. (2006) and it represents the predominant family in the order in comprising 23 123 84 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 56 Phylogram generated from maximum likelihood analysis based on combined LSU, and SSU sequence data. Related sequences are taken from from Dai et al. (2018). Eighty-five strains are included in the combined analyses which comprise a total of 3041 characters (1973 characters for LSU and 1068 characters for SSU) after alignment. Caliciopsis pinea (AFTOL-ID 1869) is used as the outgroup taxon. Single gene analyses were carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood value of - 27484.335626 is presented. The matrix had 1767 distinct alignment patterns, with 52.88% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.252599, C = 0.225029, G = 0.296253, T = 0.226118; substitution rates AC = 0.842370, AG = 2.253531, AT = 0.847156, CG = 1.108746, CT = 5.466539, GT = 1.000000; gamma distribution shape parameter a = 0.400506. Bootstrap values for maximum likelihood (ML) equal to or greater than 50 and Bayesian posterior probabilities (BYPP) equal to or greater than 0.9 are placed above the branches respectively. The newly generated sequences are indicated in bold and blue genera and 187 species (Dissanayake et al. 2016). Species of Botryosphaeriaceae have a cosmopolitan distribution on a wide range of plant hosts, as endophytes, saprobes and plant pathogens. Dothiorella Sacc., Michelia 2(6): 5 (1880) Notes: Dothiorella species are characterized by conidia that become pigmented and 1-septate, while they are still attached to the conidiogenous cells (Phillips et al. 2013). Presently, 30 species are accepted in the genus (Dissanayake et al. 2016). Wide host ranges and morphological plasticity within the genus have made it impossible to identify species based only on morphology. Phillips et al. (2008) have introduced the genus Spencermartinsia to accommodate the Dothiorella-like species but Yang et al. (2016b) have synonymized Spencermartinsia species into genus Dothiorella because they found that apiculate ascospores were not reliable to use for separating these two genera. Therefore we also have treated them as a single genus, Dothiorella. Dothiorella plurivora (Abdollahz., Javadi & A.J.L. Phillips) Tao Yang & Crous, in Yang, Groenewald, Jami, Cheewangkoon, Abdollahzadeh, Lombard & Crous 2016 Facesoffungi number: FoF05785; Fig. 57 Saprobic on dead wood. Sexual morph Undetermined. Asexual morph Coelomycetous. Conidiomata 155– 260 lm diam., stromatic, superficial, unilocular, glabrous, brown to black, globose to subglobose, papillate. Ostiole central, short, lined with periphyses. Wall of conidiomata 20–45 lm membranaceous, composed of dark brown, or brown to hyaline cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 5– 9 9 1.5–5 lm ( x = 7.5 9 3.5 lm, n = 30), enteroblastic, phialidic, cylindrical to subcylindrical, hyaline, smooth- 85 walled, arising from the inner layers of conidioma. Conidia 14–22 9 5–10 lm ( x = 18 9 8 lm, n = 50), ellipsoidal to oval, hyaline when young, becoming to brown when mature, smooth-walled, guttulate, 1-septate. Culture characteristics: Ascospores germinating on PDA (potato dextrose agar) within 2 days at 23 C. Sporulation after 1 week, colony 2 cm diam. gray from above and reverse, with white mycelia radiating outwards, with irregular form of margin, filamentous mycelium, with rough surface. Material examined: CHINA, Yunnan Province, Kunming, Songhuaba Lake; on dead wood, 3 September 2017; S.K. Huang (KUN HKAS 99572), living culture (KUMCC 18-0013). GenBank numbers: ITS: MK459467, LSU: MK459468. Notes: Spencermartinsia plurivora has been isolated from soil in Iran (Abdollahzadeh et al. 2014). Multi-gene analysis showed that Spencermartinsia to have close affinities with Dothiorella (Phillips et al. 2013; Slippers et al. 2013). Therefore, Yang et al. (2016b) placed Spencermartinsia as a synonym of Dothiorella. In phylogenetic analysis of Dothiorella, our strain clusters with the ex-type strain of Dothiorella plurivora (IRAN 1557C). Dothiorella plurivora is characterized by 7–10 9 3–5 lm conidiogenous cells and ellipsoid to ovoid, brown, 1-septate conidia (20–25 9 10–13 lm) (Abdollahzadeh et al. 2014). Morphological similarities reveal it is similar to our taxon. Dothiorella rhamni Wanas., Bulgakov, E.B.G. Jones & K.D. Hyde, in Li et al., Fungal Diversity 78: 253 (2016) Facesoffungi number: FoF01668; Fig. 58 Saprobic or weak pathogen on dead twigs of Cercis canadensis L. (Fabaceae). Sexual morph Undetermined. Asexual morph Conidiomata 310–370 lm high 9 360– 500 lm diam. ( x = 339 9 425 lm, n = 10), pycnidial, stromatic, mostly solitary, semi-immersed to immersed in the host, dark brown to black, apapillate. Peridium multilayered 50–57 lm wide at the base, 68–83 lm wide in sides, comprising 8–10 layers, heavily pigmented, thickwalled, comprising blackish to dark brown, textura angularis cells. Conidiogenous cells 9–15 lm high 9 2–4 lm wide, holoblastic, cylindrical to subcylindrical, hyaline, the first conidium produced holoblastically and subsequent conidia enteroblastically, swollen at the base, discrete, producing a single conidium at the apex. Conidia 15– 24 9 7–11 lm ( x = 20.9 9 8.9 lm, n = 30), initially hyaline, unicellular, becoming cinnamon to sepia and 1-septate, while still attached to conidiogenous cells; detached conidia, hyaline, sepia or dark brown, unicellular or 1-septate moderately thick-walled, wall externally smooth, roughened on the inner surface, oval to ovoid, widest in the centre, apex obtuse, base truncate or rounded. 123 86 Fungal Diversity (2019) 96:1–242 Fig. 57 Dothiorella plurivora (HKAS 99572, new record). a Appearance of conidiomata on host. b, c Conidiomata in vertical section. d Wall of conidioma. e–f Conidiogenous cells with developing conidia. g–i Conidia. j Germinated conidia. Scale bars: b, c = 100 lm, d = 50 lm, j = 20 lm, e–i = 10 lm Culture characteristics: Colonies on PDA, circular, fimbriate, crenate edged, initially both surfaces white and turning to grey towards centre after 4 days, fast growing, reach 70 mm diam. in 6 days at 28 C. Material examined: RUSSIA, Rostov region, Shakhty City, Alexandrovsky (Central) Park (47.7058991 N, 40.2053368E), on dead and dying twigs of Cercis canadensis L., 1 March 2016, Timur S. Bulgakov (MFLU 16–1545), living culture (MFLUCC 19-0001). GenBank numbers: ITS: MK064561, TEF1-a: MK078544. Notes: Dothiorella rhamni was introduced by Li et al. (2016). There are several records of this species from Rhamnus cathartica in Russia, Tamarix gallica and Rhamnus alaternus in Italy (Dissanayake et al. 2016; Li et al. 2016). Based on our phylogenetic analysis of combined ITS and TEF1-a sequence data of Dothiorella species (Fig. 61), our strain (MFLUCC 19-0001) clustered with the ex-type strain of D. rhamni (MFLUCC 14-0902) and the two specimens share similar morphological characters. 123 Fungal Diversity (2019) 96:1–242 87 Fig. 58 Dothiorella rhamni (MFLU 16-1545, new record). a Appearance of conidiomata on host substrate. b Vertical section of a conidioma. c Peridium of conidioma. d, e Immature and mature conidia attached to conidiogenous cells. f, g Mature conidia. h Culture characters on PDA. Scale bars: b, c = 50 lm, d = 20 lm, f = 20 lm, e, g = 10 lm Dothiorella styphnolobii Brahmanage, Bulgakov & K.D. Hyde, sp. nov. Index Fungorum number: IF555482; Facesoffungi number: FoF05095; Fig. 59 123 88 Fungal Diversity (2019) 96:1–242 Fig. 59 Dothiorella styphnolobii (MFLU 17–2256, holotype). a, b Appearance of conidiomata on Styphnolobium japonicum. c Vertical section of conidioma. d Peridium. e Developing stages of conidia on conidiogenous cells. f Immature conidium attach to conidiogenous cells. g–j Conidia. Scale bars: b = 1000 lm, c = 100 lm, d, e = 20 lm, f–j = 10 lm Etymology: The specific epithet ‘‘styphnolobii’’ refers to the host plant genus Styphnolobium. Holotype: MFLU 17-2256 Saprobic on dead twigs and branches of Styphnolobium japonicum (L.) Schott (Fabaceae). Sexual morph Undetermined. Asexual morph Conidiomata 425–500 9 160– 180 lm ( x = 460 9 170 lm, n = 30), pycnidial, stromatic, mostly solitary, semi-immersed to immersed, dark brown to black, ostiolate. Ostiole apapillate. Peridium 30–65 lm wide, comprising 6–8 layers of heavily pigmented, thickwalled, blackish to dark brown, angular cells. Conidiogenous cells 7–9 9 2–4 lm, holoblastic, conical, hyaline, swollen at the base, discrete, producing a single conidium at the apex, proliferating at the same level giving rise to periclinal thickenings. Conidia 18–22 9 8–10 lm ( x = 20 9 8 lm, n = 30), initially hyaline, unicellular, 123 Fungal Diversity (2019) 96:1–242 guttulate, becoming pale brown and 1-septate, while still attached to conidiogenous cells; detached conidia, hyaline, dark brown, unicellular or 1-septate, slightly concentric, oval to ovoid, apex obtuse, base truncate or rounded. Culture characteristics: Colonies on PDA at 25 C reaching 5 cm in 7 days, ash at first, becoming blackish when mature and reverse ash becoming blackish when mature. Material examined: RUSSIA, Republic of Crimea, Feodosia Municipality, Karadag Nature Reserve, the park of Karadag Biological Station, on dead and dying twigs and branches of Styphnolobium japonicum (L.) Schott, 21 June 2016, Timur S. Bulgakov (MFLU 17-2256, holotype), extype living culture (JZB3150013). GenBank numbers: ITS: MH880849, TEF1-a: MK069594. Notes: Our new collection shares similar morphological characters with Dothiorella, such as 1-septate conidia that are dark at an early stage of development (Phillips et al. 2005, 2013). In our phylogenetic analyses, D. styphnolobii showed close phylogenetic affinities to D. juglandis (Fig. 61). Dothiorella styphnolobii resembles D. italica in the characteristics of conidiomata and conidia, but are distinct in the shape of conidiogenous cells that are subcylindrical in D. italica and conical in D. styphnolobii (Dissanayake et al. 2017a). Dothiorella styphnolobii can also be distinguished from D. italica by the conidiomatal width (160–180 lm vs 680 lm), conidiogenous cells (7–9 9 2–4 lm vs 8–15 9 3–6 lm) and conidia (18–22 9 8–10 lm vs 28.7–43.2 9 13.2–17.5 lm) (Dissanayake et al. 2017a). However, they formed two distinct lineages in our phylogram with the maximum likelihood support of 80% ML (Fig. 61). Dothiorella symphoricarposicola W.J. Li, J.K. Liu & K.D. Hyde, Cyptogamie Mycologie 35(3): 265 (2015) Facesoffungi number: FoF04935; Fig. 60 Saprobic on dead, aerial branch of Laburnum anagyroides Medik. Sexual morph Undetermined. Asexual morph Conidiomata 0.17–0.53 mm ( x = 0.39 mm, n = 10) diam., solitary, scattered, immersed to semi-immersed, partially erumpent at maturity, pyriform, unilocular, globose, black, ostiolate. Ostiole 30–60 lm diam., single, central, with a well-developed neck, thick-walled, sometimes papillate. Peridium multi-layered, 45 lm wide at the base, 20–38 lm wide in sides, outer layer composed of 4–5 layers of thick, brown cells, inner 3–4 layers of hyaline cells of textura angularis, cells towards inner layer become paler. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 9–14 9 3–5.3 lm ( x = 11.5 9 4.2 lm, n = 10), phialidic, cylindrical, sometimes slightly curved, hyaline, smooth, thick-walled, formed from the inner most layer of pycnidial wall. 89 Conidia 14–21 9 7–11.5 lm ( x = 18.4 9 9 lm, n = 30), ovoid, straight or slightly curved, guttulate, initially aseptate, hyaline, becoming 1-septate, pale to dark brown at maturity, smooth-walled, rounded at both ends. Culture characteristics: Colonies on PDA, circular, fimbriate, crenate edged, with both surfaces white at first, becoming grey towards centre and finally becoming black, fast growing, and reach 6 cm diam. after 4 days at 28 C. Material examined: ITALY, Forlı̀-Cesena Province, Corniolo – Santa Sofia, on dead, aerial branch of Laburnum anagyroides Medik. (Fabaceae), 15 May 2016, Erio Camporesi, IT-2971 (MFLU 16-1355), living culture (MFLUCC 18-0092, KUMCC 18-0414). GenBank numbers: ITS: MH819800, TEF1-a: MH853685. Notes: Dothiorella symphoricarposicola was introduced by Li et al. (2014). There are several records of this species from Symphoricarpos sp., Sambucus nigra, Laurus nobilis and Laburnum alpinum in Italy (Li et al. 2014; Dissanayake et al. 2016). Here we introduce D. symphoricarposicola as a new host record. Based on our phylogenetic analysis of combined ITS and TEF1-a sequence data of Botryosphaeriaceae species (Fig. 61), our strain (MFLUCC 18-0092) clusters with the ex-type strain of D. symphoricarposicola (MFLUCC 13-0497). The two strains share similar morphological characters. However, our strain has slightly larger conidiomata (0.17–0.53 mm) and conidiogenous cells (9–14 9 3–5.3 lm) compared to the conidiomata (0.2–0.3 mm) and conidiogenous cells (4–12 9 1.5–6 lm) of the ex-type strain (Li et al. 2014). Lasiodiplodia Ellis & Everh. Notes: This genus comprises 53 species (Dissanayake et al. 2017a). Both sexual and asexual morphs have been reported within the genus (Alves et al. 2008; Tennakoon et al. 2016b). For the species differentiation morphology is not a reliable character and species can be resolved using combined ITS and TEF1-a sequence data (Phillips et al. 2013; Slippers et al. 2014). Lasiodiplodia iraniensis Abdollahz., Zare & A.J.L. Phillips [as ‘iranensis’], Persoonia 25: 8 (2010) Facesoffungi number: FoF04922; Fig. 62 Saprobic on dried pods of Cassia sp. Sexual morph Undetermined. Asexual morph Appearing as raised spots on the host. Coelomycetous. Conidiomata 210–405 lm high 9 90–336 lm diam. ( x = 283 9 194 lm, n = 10), stromatic, pycnidial, immersed, aggregated, unilocular, globose to depressed globose. Peridium 20–55 lm wide, composed of 3–6-layers of pale brown to brown, smoothwalled cells of textura angularis. Pseudoparaphyses 2.1– 3.9 lm wide, aseptate, hyaline. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 10–22 lm 9 3– 6.3 lm ( x = 13 9 4.4 lm, n = 20), lining the pycnidial 123 90 Fungal Diversity (2019) 96:1–242 Fig. 60 Dothiorella symphoricarposicola (MFLUCC 18-0092, new host record). a Appearance of conidiomata on dead branch of Laburnum anagyroides. b Longitudinal section of a conidioma. c Longitudinal section of a conidioma wall showing cell organization. d–f Conidiogenous cells with developing conidia. g Conidia. h, i Colony on PDA (h from below view, i from above view). Scale bars: a = 0.5 mm, b = 100 lm, c–g = 20 lm cavity, holoblastic, subcylindrical, hyaline, smooth-walled. Conidia 14–28 9 8–14.5 lm ( x = 22.5 9 12, n = 25) ovoid, rarely subovoid to ellipsoid-ovoid, with a broadly rounded apex and truncate base, hyaline, thick-walled, smooth-walled. Culture characteristics: Colonies on PDA reaching 55 mm diam. after 7 days at 25 C, circular, smooth 123 Fungal Diversity (2019) 96:1–242 margin, white at first, ash to greenish black after 2 weeks, flat on the surface, with aerial mycelium, reverse black. Material examined: THAILAND, Chiang Rai Province, on dried pods of Cassia sp. (Fabaceae), 22 December 2016, R.H. Perera, RHP 37 (MFLU 17-2847), living culture (MFLUCC 18-0586). GenBank numbers: ITS: MH107831, LSU: MH107832, TEF1-a: MH107833, TUB2: MH107834. Notes: The new collection is a typical Lasiodiplodia species, and it shares the similar morphology with Lasiodiplodia iraniensis which the dimensions of their conidiophores and conidia are almost identical (Abdollahzadeh et al. 2010). Conidia of L. iraniensis become 1-septate and dark brown with longitudinal striations with the time (Abdollahzadeh et al. 2010). However, we did not observe released mature conidia in our collection. The multi-gene phylogenetic analysis showed that our newly obtained strain (MFLUCC 18-0586) clustered together with Lasiodiplodia iraniensis (Fig. 63). Therefore, we identify our collection as L. iraniensis and it is reported here as a new record to Thailand and Cassia sp. Jahnulales K.L. Pang, Abdel-Wahab, El-Shar., E.B.G. Jones & Sivichai Notes: Jahnulales was erected by Pang et al. (2002) which comprises freshwater lignicolous ascomycetes (Huang et al. 2018). This order is phylogenetically related to the Dothideales, Patellariales, and Pleosporales (Campbell et al. 2007). Jahnulales species often occur on rotting or soft submerged corticated or decorticated wood (Inderbitzin et al. 2001; Suetrong et al. 2011a; Tanaka et al. 2015; Huang et al. 2018). They produce ascomata with multilayered peridial walls, composed of large cells, stalked and/ or sessile bitunicate asci, and one-septate ascospores with appendages or gelatinous sheaths (Pang et al. 2002; Suetrong et al. 2011a; Jones et al. 2015). Two families, Aliquandostipitaceae and Manglicolaceae are currently accepted within this order (Jones et al. 2015; Wijayawardene et al. 2018a). Aliquandostipitaceae Inderb. Notes: Aliquandostipitaceae, typified by Aliquandostipite Inderb. is characterized by globose to subglobose ascomata, and one-septate ascospores (Inderbitzin et al. 2001). The sexual morph genera, Aliquandostipite, Jahnula and Megalohypha (Kirschstein 1936; Inderbitzin et al. 2001; Ferrer et al. 2007), and asexual morph genera, Brachiosphaera, Speiropsis and Xylomyces are currently accommodated within this family (Tubaki 1958; Descals et al. 1976; Goos et al. 1977; Campbell et al. 2007; Suetrong et al. 2011a; Wijayawardene et al. 2018a). 91 Jahnula Kirschst. Notes: Jahnula was introduced by Kirschstein (1936) and typified by Jahnula aquatica (Kirschst.) Kirschst. Jahnula species have been reported from wood or decorticated wood in freshwater habitats (Hawksworth 1984; Hyde 1993; Hyde and Goh 1998; Ho et al. 2002; Suetrong et al. 2011a; Fournier et al. 2015; Huang et al. 2018). The genus is polyphyletic, however, Jahnula sensu stricto accommodates J. aquatica, J. granulosa, and J. rostrata (Suetrong et al. 2011a). We introduce an additional taxon to this genus from decaying wood submerged in a river in Australia. Jahnula queenslandica Dayarathne, Fryar & K.D. Hyde, sp. nov. Index Fungorum number: IF555384; Facesoffungi number: FoF04945; Fig. 64 Etymology: The name queenslandica refers to the geographic location where the specimen was collected. Holotype: AQ522780 Saprobic on dead wood. Sexual morph Ascomata 280– 330 9 200–300 lm ( x = 350 9 250 lm, n = 5), perithecial, solitary, superficial to semi-immersed, unilocular, obpyriform to subglobose, dark brown to black, papillate, ostiolate. Ostiole central, composed of hyaline periphyses. Peridium 28–40 lm thick, membranous, composed of brown cells of textura angularis. Hamathecium comprising 1.5–2 lm wide, septate, branched, filiform pseudoparaphyses, embedded in a gelatinous matrix. Asci 140– 200 9 12–22 lm ( x = 180 9 16 lm, n = 20), 8-spored, bitunicate, fissitunicate, cylindrical, pedicellate, rounded at apex, with a distinct ocular chamber. Ascospores 20– 34 9 8–16 lm ( x = 25 9 12 lm, n = 30), uniseriate, initially light brown, becoming dark brown at maturity, oval to broadly ellipsoid, aseptate, with a mammiform apex, slightly curved, smooth-walled, often with two large guttules, apiosporous, rounded at lower end, with hyaline cell at lower end, up to 2–3 lm, lacking a sheath. Asexual morph undetermined. Material examined: AUSTRALIA, Queensland, Daintree National Park, fast running stream with clear water and a rocky base, S 16.170779, E 145.405268, on decaying wood submerged in a river, 14 April 2015, Sally Fryar and Toby Cawson, CT110 (AQ522780, holotype; MFLU 18-1692, isotype). GenBank numbers: LSU: MH878780, ITS: MH878782. Notes: Jahnula queenslandica has obpyriform to subglobose ascomata, septate, branched, filiform pseudoparaphyses, cylindrical asci and oval to broadly ellipsoid ascospores that are characteristic of Jahnula (Raja et al. 2008; Sivichai and Boonyuen 2010; Suetrong et al. 2011a; 123 92 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 61 Phylogram generated from maximum parsimony analysis based on combined ITS and TEF1-a sequence data for Dothiorella species in family Botryosphaeriaceae. Related sequences are taken from Wanasinghe et al. (2018a) and GenBank. Sixty-six strains are included in the combined analyses which comprise 828 characters after alignment. Lasiodiplodia theobromae (CBS 164.96) (Botryosphaeriaceae and Botryosphaeriales) is used as the outgroup taxon. Values for maximum likelihood equal to or greater than 60 and Bayesian posterior probabilities equal or greater than 0.90 are placed. Maximum parsimony analysis of 85 parsimony informative characters resulted in a most parsimonious tree (CI = 0.545, RI = 0.824, RC = 0.449, HI = 0.455). The ex-type strains are in bold and black. The newly generated sequences are in bold and blue Fournier et al. 2015; Huang et al. 2018). However, J. queenslandica is unique from other Jahnula species by having apiospores. According to our phylogram (Fig. 65) J. queenslandica (MFLU 18-1692) is closely related to J. sunyatsenii (UBC F13876) (Inderb.) K.L. Pang, E.B.G. Jones & Sivichai with high statistical support (91% ML/ 0.99 BYPP). Jahnula queenslandica can be distinguish from J. sunyatsenii by its aseptate, dark brown apiospores, while ascospores of J. sunyatsenii are 1-septate, light brown with two bipolar, helmet-shaped appendages, tending to unite over the respective ends (Raja and Shearer 2006). Tubeufiales Boonmee & K.D. Hyde Notes: The order Tubeufiales was introduced by Boonmee et al. (2014) for a single family, Tubeufiaceae. Later, Suetrong et al. (2014) introduced the second family Wiesneriomycetaceae into Tubeufiales. Recently, based on divergence time estimates, Liu et al. (2017) synonymized the orders Bezerromycetales and Wiesneriomycetales under Tubeufiales. Tubeufiaceae M.E. Barr Notes: Barr (1979) introduced the family Tubeufiaceae typified by Tubeufia. According to recent study by Lu et al. (2018b) this family comprises 38 genera. The sexual morphs of Tubeufiaceae are characterized by having superficial ascomata, pseudoparaphysate hamathecium, bitunicate asci, and multi-septate, hyaline to pale brown cylindrical ascospores (Barr 1980; Lu et al. 2017; Liu et al. 2018; Brahmanage et al. 2017; Phookamsak et al. 2018). Hyphomycetous asexual morphs of Tubeufiaceae are often helicosporous, while some are chlamydosporous and phragmosporous (Boonmee et al. 2011, 2014; Brahmanage et al. 2017; Doilom et al. 2017; Luo et al. 2017; Liu et al. 2018; Lu 2018). Kamalomyces R.K. Verma, N. Sharma & Soni Notes: The genus Kamalomyces introduced by Verma et al. (2008), is typified by K. indicus R.K. Verma, N. Sharma & Soni. Kamalomyces has a unique set of 93 characters, such as ascomata on a subiculum of black hyphae, with solitary, gregarious, subglobose to lemoniform ascomata with short stalks and lacking ostioles, bitunicate, broadly cylindrical to clavate asci and hyaline, vermiform, crowded, septate ascospores (Verma et al. 2008; Dubey and Neelima 2013; Boonmee et al. 2011, 2014; Phookamsak et al. 2018; Lu et al. 2018a, b). Kamalomyces currently comprises four species viz. K. bambusicola Phook., Y.Z. Lu & K.D. Hyde, K. indicus, K. mahabaleshwarensis Rashmi Dubey & Moonamb. and K. thailandicus Phook., Y.Z. Lu & K.D. Hyde (Verma et al. 2008; Dubey and Neelima 2013; Phookamsak et al. 2018). The phylogenetic relationships of Kamalomyces were reported by Phookamsak et al. (2018). We introduce a novel marine species to this genus based on combined LSU, ITS and TEF1-a sequence data. Kamalomyces mangrovei Dayarathne & K.D. Hyde, sp. nov. Index Fungorum number: IF555385; Facesoffungi number: FoF04946; Fig. 66 Etymology: Epithet derived from the mangrove habitat that species found. Holotype: MFLU 18-1691 Saprobic on decaying, submerged wood in a mangrove stand. Sexual morph Ascomata 280–315 9 250–300 lm, superficial, solitary to gregarious, embedded in a subiculum of crowded, black, septate, thick-walled hyphae, superficial, solitary, gregarious, globose to subglobose, glabrous, shortstalked, apapillate, with indistinct ostiolate. Peridium 30– 45 lm wide, comprising light brown cells of textura angularis, and inwardly small, subhyaline cells of textura prismatica. Hamathecium comprising numerous, 1.5–2 lm wide, filiform, septate, branched, hyaline pseudoparaphyses. Asci 148–180 9 16–20 lm ( x = 165 9 18 lm, n = 20), 8spored, bitunicate, fissitunicate, cylindrical to clavate, short pedicellate, apically rounded, with an ocular chamber. Ascospores 52–67 9 6.5–8.5 lm ( x = 62 9 7.5 lm, n = 30), 2–3-seriate, hyaline becoming light brown when mature, elongate cylindrical to fusiform-clavate, tapering towards the lower cells, enlarged at the 4th and 5th cell, straight or slightly curved, 8–9 septa, distoseptate. Culture characteristics: Colonies on PDA reaching 3 cm diam. after 30 days at 25 C, circular, smooth margin white at first, dark gray to black after 6 weeks, flat on the surface, without aerial mycelium, reverse brownish black. Material examined: THAILAND, Ranong Province, Maung District, Mu 4 Tombol Ngao, Ranong Mangrove Research Center (GPS: 9430 to 9570 N; 98290 to 98390 E) on decaying, submerged wood of mangrove, 7 December 2016, Monika C. Dayarathne, MCD 053 (MFLU 18-1691, holotype), ex-type living culture (MFLUCC 17-0407, TBRC). 123 94 Fungal Diversity (2019) 96:1–242 Fig. 62 Lasiodiplodia iraniensis (MFLU 17-2847, new host/geographical record). a Herbarium material. b Conidiomata on host substrate. c, d Sections through conidiomata. e Peridium. f Paraphyses in Congo red. g, h Conidiophores with conidia. i–k Conidia. l Germinating conidium. m, n Colony on PDA. Scale bars: c, d = 200 lm, e = 50 lm, f–i = 20 lm GenBank numbers: ITS: MH878781, LSU: MH878779, SSU: MH878796, TEF1-a: MH886508. Notes: Kamalomyces mangrovei is characterized by solitary to gregarious, globose to subglobose, short-stalked ascomata, lacking ostioles embedded in a subiculum of black hyphae, bitunicate, cylindrical to clavate asci and hyaline, elongate cylindrical to fusiform-clavate, septate ascospores. Kamalomyces mangrovei is morphologically best fit Kamalomyces (Verma et al. 2008; Dubey and Neelima 2013; Boonmee et al. 2011, 2014, Phookamsak et al. 2018; Lu et al. 2018b). Kamalomyces mangrovei closely resembles K. bambusicola Y.Z. Lu & K.D. Hyde and K. thailandicus Phook., Y.Z. Lu & K.D. Hyde due to its ascomatal characters, asci and ascospore shape. However, K. mangrovei differs significantly from K. bambusicola and K. thailandicus by having fewer ascospore septa (8–9 vs 27–30 and 33–36 septa) and being swollen at the 4th and 5th cells. The novel species occurred on a mangrove species while K. bambusicola and K. thailandicus have been reported from bamboo. This is the first record of 123 Fungal Diversity (2019) 96:1–242 Tubeufiaceae from a marine habitat (Jones et al. 2015, 2019). According to our phylogenetic analyses with concatenated LSU, ITS and TEF1-a sequence data, K. mangrovei clustered within the genus Kamalomyces with high statistical support (84% ML/75% MP/0.99 BYPP, Fig. 67). Eurotiomycetes O.E. Erikss. & Winka Chaetothyriomycetidae Doweld Chaetothyriales M.E. Barr Notes: Barr (1987) established the order Chaetothyriales based on the presence of periphysoids in the ascomata. However, Loculoascomycetes was suggested to be paraphyletic and Chaetothyriales had closer affinity with Eurotiales based on SSU sequence data (Haase et al. 1995; Spatafora et al. 1995; Berbee 1996; Silva-Hanlin and Hanlin 1999). Doweld (2001) introduced the subclass Chaetothyriomycetidae for Chaetothyriales. Chaetothyriales comprises ecologically diverse species including opportunistic pathogens on humans (Réblová et al. 2013; de Hoog 2014; Hyde et al. 2018a, b). Five families are currently recognized in Chaetothyriales, viz. Chaetothyriaceae, Cyphellophoraceae, Epibryaceae, Herpotrichiellaceae, and Trichomeriaceae (Teixeira et al. 2017). Herpotrichiellaceae Munk Notes: The family Herpotrichiellaceae was introduced by Munk (1953), based on the type genus Herpotrichiella Petr., along with Berlesiella Sacc., Capronia Sacc., Dictyotrichiella Munk and Didymotrichiella Munk. Sexual morphs in this family are characterized by ascomata with setae and ostioles, bitunicate, saccate to ovoid asci with a thickened apex and pale grey to brown ascospores (Munk 1953), while asexual morphs are diverse dematiaceous hyphomycetes (Gueidan et al. 2014; Klaubauf et al. 2014; Liu et al. 2015; Tian et al. 2016; Dong et al. 2018). However, the classification of some hyphomycetes in this family has been controversial. Some previously morphologically-based species were excluded from Herpotrichiellaceae with molecular evidence. For example, Veronaea simplex Papendorf was transferred to Venturiaceae (Arzanlou et al. 2007). Some genera, e.g. Cladophialophora and Exophiala require further study to confirm their natural classification (Liu et al. 2015; Hyde et al. 2016; Tian et al. 2016; Dong et al. 2018). Currently, Herpotrichiellaceae comprises 15 genera and resides in the order Chaetothyriales (Liu et al. 2015; Hyde et al. 2016; Wijayawardene et al. 2018a). Thysanorea Arzanlou Notes: Arzanlou et al. (2007) studied ramichloridiumlike species and revealed that the genus Periconiella was polyphyletic. Thysanorea was therefore introduced to accommodate Periconiella papuana Aptroot as T. papuana (Aptroot) Arzanlou, W. Gams & Crous (Arzanlou et al. 95 2007). Thysanorea differs from Periconiella in branching pattern of conidiophores and prominent denticle-like conidiogenous loci (Arzanlou et al. 2007). Dong et al. (2018) introduced the second species, T. aquatica W. Dong, H. Zhang & K.D. Hyde from an aquatic habitat. Only two species were accepted in Thysanorea (Arzanlou et al. 2007; Dong et al. 2018). Thysanorea uniseptata N.G. Liu & K.D. Hyde, sp. nov. Index Fungorum number: IF555387; Facesoffungi number: FoF04949; Fig. 68 Etymology: Named after the fact that conidia are 1-septate. Holotype: MFLU 18-1723 Saprobic on submerged wood. Sexual morph Undetermined. Asexual morph Colonies on natural substrate effuse, scattered, velvety, dark brown to black, with greyish white, glistening heads of conidia at apex. Mycelium partly immersed, partly superficial, composed of pale brown to brown, branched, septate, smooth hyphae. Conidiophores up to 310 lm long, 5.5–8 lm wide at base, 3– 5 lm wide in upper half, macronematous, mononematous, branched at apex, erect, straight or broadly curved, septate, constricted at septa near apex, not constricted at septa below, thick-walled, smooth-walled, brown below, paler towards apex. Conidiogenous cells 6–11.5 9 2–4 lm ( x = 8.9 9 3.2 lm, n = 15), polyblastic, terminal or intercalary at main stem and fertile branches, hyaline or subhyaline, urniform, conidiogenous loci obvious. Conidia 6.5–9 lm long, 2–3 lm wide at the septum ( x = 7.6 9 2.5 lm, n = 30), acropleurogenous, solitary, pyriform, hyaline and aseptate when young, subhyaline, (0–)1-septate when mature, not constricted at the septa, broadly round at the apex, attenuate and narrowly truncate at the base, with darkened hilum, smooth-walled. Culture characteristics: Conidia germinated on water agar media within 24 h. Germ tubes produced from one or both ends of conidium. Colonies growing on PDA circular, edge entire, with umbonate surface, greyish brown to brown from above, dark brown from reverse, mycelium dense, fluffy. Material examined: THAILAND, Chiang Rai Province, Muang District, Ban Nang Lae Nai, on decaying wood submerged in a freshwater stream, 6 March 2018, N.G. Liu, CR067 (MFLU 18-1723, holotype), ex-type living culture (MFLUCC 18-0701). GenBank numbers: ITS: MH883033, LSU: MH883031, SSU: MH883049. Notes: Thysanorea uniseptata resembles T. aquatica and T. papuana in having obvious denticle-like conidiogenous loci and pyriform, (0–)1-septate conidia. However, conidia of T. aquatica have prominent guttules and are constricted at septa, features that distinguish it from T. uniseptata. 123 96 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 63 Phylogram generated from maximum parsimony analysis based on combined ITS and TEF1-a sequence data for Lasiodiplodia species in family Botryosphaeriaceae. Related sequences are taken from Wanasinghe et al. (2018a). Sixty-three strains are included in the combined analyses which comprise 1300 characters after alignment. Neofusicoccum luteum (CBS 110497) (Botryosphaeriaceae and Botryosphaeriales) is used as the outgroup taxon. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. Maximum parsimony analysis of 575 constant characters and 79 informative characters resulted in most parsimonious tree (CI = 0.551, RI = 0.769, RC = 0.423, HI = 0.449). The best RaxML tree with a final likelihood value of - 4380.160038 is presented. The matrix had 317 distinct alignment patterns, with 36.03% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.215525, C = 0.287133, G = 0.258862, T = 0.238480; substitution rates AC = 0.872639, AG = 3.555191, AT = 1.330285, CG = 1.135935, CT = 5.468515, GT = 1.000000; gamma distribution shape parameter a = 0.196200. Bootstrap values for maximum parsimony (MP) and maximum likelihood (ML) equal to or greater than 65 and Bayesian posterior probabilities (BYPP) equal or greater than 0.90 are placed below the branches respectively. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Thysanorea uniseptata differs from T. papuana in having hyaline or subhyaline conidiogenous cells, while the latter are golden to dark brown. Moreover, T. uniseptata has smaller conidiogenous cells and larger conidia than those of T. papuana (6–11.5 vs 5–20 lm long and 6.5–9 vs 5–6 lm long). Phylogenetic analysis showed that T. uniseptata is phylogenetically distinct from T. aquatica and T. papuana (Fig. 69). Eurotiomycetidae Geiser & Lutzoni Eurotiales G.W. Martin ex Benny& Kimbr. Notes: Currently, the order Eurotiales comprise three families, Aspergillaceae Link, Thermoascaceae Apinis and Trichocomaceae E. Fisch (Houbraken and Samson 2011). Some species in the Eurotiales are able to grow in extreme environments, such as low water activity, low acidity, low levels of oxygen, or at extreme high temperatures (Houbraken and Samson 2011; Houbraken et al. 2014). Aspergillaceae Link Notes: The family Aspergillaceae was erected by Link (1826) and belongs to Eurotiales. The most well-known species of this family belongs to genera Aspergillus P. Micheli ex Haller and Penicillium Link (Houbraken and Samson 2011; Houbraken et al. 2014). Several species belong to Aspergillaceae have the ability to produce a diversity of secondary metabolites (or extrolites), organic acids and diverse enzymes that degrade a wide variety of complex biomolecules (Geiser et al. 2006; Pitt and Hocking 2009; Samson et al. 2010; Houbraken and Samson 2011), while others can cause disease in humans. 97 Penicillium Link Notes: The genus is subdivided in two subgenera (Aspergilloides and Penicillium) and 26 sections (Houbraken and Samson 2011; Houbraken et al. 2016). Species of Penicillium are ubiquitous and known to cause rot in fruits and bulbs of plants, including P. expansum (apples and pears), P. digitatum, P. italicum and P. ulaiense (citrus fruits), P. tulipae (tulip bulbs) and P. allii (garlic) (Balgrie 2003; Samson and Frisvad 2004; Valdez et al. 2006). Currently, the genus contains more than 350 accepted species (Visagie et al. 2014). Section Fasciculata contains species that are found on stored or manufactured foods, and have rough-walled conidiophore stipes and subglobose conidia (Houbraken et al. 2016). Species in this section are able to grow at low water activities, and low temperature (with the exception of species in series Verrucosa) (Houbraken et al. 2016). Section Fasciculata is subdivided in four series: Camemberti, Corymbifera, Viridicata, Verrucosa. Six species are currently accepted in series Corymbifera (P. albocoremium, P. allii, P. hirsutum, P. hordei, P. radicicola, P. tulipae) and most of those species (except P. hordei) have a strong association with flower bulbs and other plant roots (Houbraken et al. 2016). During an investigation of the fungal diversity on roots of deodeok (Codonopsis lanceolata), rotten apple (Malus pumila), freshwater, and soil in Korea, five new species of Penicillium were isolated and these are described here based on phylogenetic analyses (Figs. 70, 71, 72, 73) and morphological characteristics (Figs. 74, 75, 76, 77, 78). Penicillium circulare Hyang B. Lee, P.M. Kirk & T.T.T. Nguyen sp. nov. Index Fungorum number: IF555413; Facesoffungi number: FoF05793; Fig. 74 Etymology: circulare, referring to the formation of two concentric rings on YES media. Holotype: CNUFC-GEU220-1 Macromorphology: CYA, 25 C: Sporulation on CYA moderate; green in sporulating area with white mycelium in centre, yellow at the margins; radial wrinkles; exudate absent; soluble pigment absent; reverse orange to dark orange brown, wrinkled towards the edges; YES, 25 C: Weak sporulation on YES; exudate absent; soluble pigments absent; reverse buff, two concentric rings. MEA, 25 C: Sporulation on MEA strong; mycelium white turns pale yellow at the margins; exudate absent; soluble pigments absent; reverse pale brown or orange brown towards the centre. Micromorphology: Sclerotia absent. Conidiophores monoverticillate, stipe smooth, 2.5–3.5 lm wide, unbranched or with one or two branches 25–58 lm long. 123 98 Fig. 64 Jahnula queenslandica (MFLU 18-1692, holotype). a Host. b Appearance of ascomata on host. c Section of ascoma. d Periphysate ostioler region. e Peridium. f, g Asci. h Apex of an ascus. 123 Fungal Diversity (2019) 96:1–242 i Pseudoparaphyses. j, k Apiosporous ascospores. Scale bars: c, f, g, i = 100 lm, d = 50 lm, e = 20 lm, h, j, k = 10 lm Fungal Diversity (2019) 96:1–242 99 Fig. 65 Phylogram generated from maximum likelihood analysis based on combined LSU and ITS sequence data of selected taxa. Related sequences are obtained from GenBank. Forty-nine strains are included in the analyses, which comprise 1159 characters including gaps. Single gene analyses are carried out and compared with each species, to compare the topology of the tree and clade stability. The tree is rooted with Farlowiella carmichaeliana (JX 43). Tree topology of the ML analysis is similar to the MP and BI. The best scoring RAxML tree with a final likelihood value of - 9016.535047 is presented. The matrix had 607 distinct alignment patterns, with 27.68% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.244145, C = 0.229207, G = 0.302247, T = 0.224401; substitution rates AC = 0.875220, AG = 1.547245, AT = 1.078795, CG = 0.865551, CT = 4.551536, GT = 1.000000; gamma distribution shape parameter a = 0.372555. The maximum parsimonious dataset consisted of constant 604, 385 parsimonyinformative and 170 parsimony-uninformative characters. The parsimony analysis of the data matrix resulted in the maximum of two equally most parsimonious trees with a length of 1569 steps (CI = 0.556, RI = 0.754, RC = 0.420, HI = 0.444) in the first tree Maximum parsimony bootstrap (MPBT, blue) values [ 65%, Bayesian posterior probabilities (PP, green) [ 0.80% and maximum likelihood bootstrap (ML, black) values [ 65%) are given above the nodes. The scale bar indicates 0.07 changes. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Phialides ampulliform to cylindrical shaped, 3–10 per metula, 7.5–11.5 9 2.5–3.5 lm. Conidia globose 2.5–3.5 lm. Culture characteristics: The isolate grew over a wide range of temperatures with varying growth rates on MEA, CYA, and YES. The average growth rates of CNUFCGEU220-1 on MEA, CYA, and YES medium at 25 C 123 100 Fig. 66 Kamalomyces mangrovei (MFLU 18-1691, holotype). a, b Ascomata. c. Section of ascoma. d Section through peridium. e–g Asci and pseudoparaphyses. h–k Ascospores. l Germinating 123 Fungal Diversity (2019) 96:1–242 ascospore. m, n Cultures on PDA (m from above view, n from below view). Scale bars: a = 500 lm, c = 200 lm, b = 100 lm, e– g = 50 lm, d, h–l = 20 lm Fungal Diversity (2019) 96:1–242 were 12, 26, and 10.5 mm per 7 days, respectively. Optimal growth was observed around 25 C, slow growth was observed at below 10 C, and no growth at 37 C. Material examined: REPUBLIC OF KOREA, Geumsan Park, Jeju Island, from forest soil, 27 March 2018 (CNUFC-GEU220-1, holotype), ex-type living culture preserved as glycerol stock at Rhizophydium koreanum 80 C in the CNUFC. GenBank numbers: CAL: MK481061, RPB2: MK481053, MK481054, TUB2: MK481057, MK481058. Notes: Penicillium circulare belongs to section Sclerotiora and is phylogenetically related to P. jacksonii, but differs by the production of two concentric rings on YES. In addition, it produces monoverticillate conidiophores with one or two branches, in contrast to a single branch of P. jacksonii. Penicillium geumsanense Hyang B. Lee, P.M. Kirk & T.T.T. Nguyen sp. nov. Index Fungorum number: IF555412; Facesoffungi number: FoF05794; Fig. 75 Etymology: geumsanense, referring to its collection location, Geumsan Park, Jeju Island from where the species was isolated (Korea). Holotype: CNUFC-GEU2229-1 Macromorphology: CYA, 25 C: Sporulation on CYA moderate; dark grey green; velvety floccose in centre; mycelium white; soluble pigment absent; reverse orange to dark orange brown; MEA, 25 C: Colonies grey green; velvety; floccose in centre; strongly fasciculate; soluble pigment absent; reverse dark orange brown and light yellow; YES, 25 C: Sporulation on YES strong; mycelium white, radially sulcate; soluble pigments absent; reverse white to dark orange brown. Micromorphology: Sclerotia absent. Conidiophores mostly terverticillate, sometimes mono-or biverticillate, occasionally with a divaricate branch, branches 22 lm long, stipe 3.0–4.5 lm wide. Metulae mostly 2–3, 9.0–17.5 9 2–4 lm. Phialides ampulliform shaped, 1–5 per metula, 7.5–10.5 9 2–3.5 lm. Conidia ellipsoidal, 3.5–4.0 9 3.0–3.5 lm. Culture characteristics: The isolate was observed to grow over a wide range of temperatures with varying growth rates on MEA, CYA, and YES. The average growth rates of CNUFC-GEU2229-1 on MEA, CYA, and YES medium at 25 C were 6.5, 11.5, and 13.5 mm per 7 days, respectively. Optimal growth was observed around 25 C, slow growth was observed at below 10 C, and no growth at 37 C. Material examined: REPUBLIC OF KOREA, Geumsan Park, Jeju Island, from forest soil, 27 March 2018 101 (CNUFC-GEU2229-1, holotype), ex-type living culture preserved as glycerol stock at - 80 C in the CNUFC. GenBank numbers: CAL: MK481062, MK481063, RPB2: MK481055, MK481056, TUB2: MK481059, MK481060. Notes: Penicillium geumsanense belongs to section Robsamsonia and is phylogenetically related to P. glandicola, a new species described below. It differs from P. glandicola by producing terverticillate, monoverticillate or biverticillate conidiophores, in contrast to the terverticillate conidiophores of P. glandicola, in addition, grows more slowly than P. glandicola on all media. Penicillium mali-pumilae Hyang B. Lee, T.T.T. Nguyen & Houbraken, sp. nov. Index Fungorum number: IF555410; Facesoffungi number: FoF05790; Fig. 76 Etymology: mali-pumilae, referring to Malus pumila, the type was isolated from rotten fruit of apple. Holotype: CBS H-22503 Macromorphology: CYA, 25 C: Colonies plane; sporulation on CYA strong; colony texture fasiculate; mycelium white; exudate present as large clear or pale yellow droplets; soluble pigment absent; radial sulcate; margin entire; conidia green to dull blue green; reverse pale brown with (dark) brown in centre. YES, 25 C: Sporulation absent, mycelium (pale) yellow in centre, white near edge; exudate present as yellow pigments; soluble pigments absent; reverse pale brown, sometimes with brown centre. MEA, 25 C: Sporulation on MEA strong; colony texture fasiculate; mycelium white; exudate present as large pale yellow or brown droplets; soluble pigments absent; conidia green occasionally blue-green; reverse yellow with pale brown centre or pale brown with brown centre. DG18, 25 C: Sporulation absent or poor; texture fasiculate; mycelium white or pale yellow, conidia dull green; reverse (pale) yellow. Micromorphology: Sclerotia absent. Conidiophores mostly long, with rough walled stipes, non-vesiculate, predominantly terverticillate, branches 20 lm long, stipe 2.5–3.5 lm wide. Metulae 4–10, 9–13 9 2.5–4 lm. Phialides flask shaped, 3–8 per metula, 8–12 9 2.5–3.5 lm. Conidia in long, distorted chains, smooth walled or very finely roughened, globose to subglobose, 3–3.5 lm. Colony diameters: 7 d, in mm: CYA 22–32; CYA15 C 26–31; CYA30 C 4–15; CYA37 C no growth; DG18 21–31; MEA 19–28; YES 21–32; CYAS 20–28; creatine agar 10–20, weak growth, poor acid production, base formation absent. Material examined: NETHERLANDS, Lilium bulb, Aalsmeer, CBS 127.90 = IBT 12175; REPUBLIC OF KOREA, rot apple (Malus pumila), stored in refrigerator 123 102 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 67 Phylogram generated from maximum likelihood analysis based on combined LSU, ITS and TEF1-a sequence data of selected taxa. Related sequences are obtained from GenBank. Data set comprises 2530 characters including gaps. Single gene analyses are carried out and compared with each species, to compare the topology of the tree and clade stability. Tree is rooted with Botryosphaeria agaves (MFLUCC 10-0051) and B. dothidea (CBS 115476). Tree topology of the ML analysis is similar to the MP and BI. The best scoring RAxML tree with a final likelihood value of - 27950.184903 is presented. The matrix had 1137 distinct alignment patterns, with 26.87% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.234676, C = 0.258513, G = 0.266669, T = 0.240142; substitution rates AC = 1.053687, AG = 4.072971, AT = 3.048439, CG = 0.739670, CT = 8.539737, GT = 1.000000; gamma distribution shape parameter a = 0.712704. The maximum parsimonious dataset consisted of constant 1587, 768 parsimonyinformative and 175 parsimony-uninformative characters. The parsimony analysis of the data matrix resulted in the maximum of two equally most parsimonious trees with a length of 5615 steps (CI = 0.273, RI = 0.584, RC = 0.160, HI = 0.727) in the first tree. Maximum likelihood bootstrap (ML, left), maximum parsimony bootstrap (MP, middle) values [ 65%, and Bayesian posterior probabilities (BYPP, right) [ 0.80% are given above the nodes. The scale bar indicates 0.05 changes. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue (4 C), Yesan, 2012, CBS 140671 = DTO 327-D1 = EML-MP6080-1. GenBank numbers: CAL: KP900992, ITS: KP900991, RPB2: MK033522, TUB2: KP900993. Notes: Penicillium mali-pumilae resembles P. psychrotrophicum and detailed notes are given in the description of that species. Penicillium psychrotrophicum Hyang B. Lee, H.Y. Mun, J.C. Frisvad & Houbraken, sp. nov. Index Fungorum number: IF555409; Facesoffungi number: FoF05789; Fig. 77 Etymology: psychrotrophicum, refers to its ability to grow at low temperature conditions. Holotype: CBS H-22504 Macromorphology: CYA, 25 C: Colonies plane; sporulation on CYA strong; colony texture fasiculate; mycelium white; exudate present as large clear droplets; soluble pigment absent; radial sulcate; margin entire; conidia green; reverse pale brown, brown in centre. YES, 25 C: Sporulation poor, mycelium white; exudate absent; soluble pigments absent; conidia dull green; reverse brown. MEA, 25 C: Sporulation on MEA strong; colony texture fasiculate; mycelium white; exudate present as large clear droplets; soluble pigments absent; conidia green; reverse yellow with pale brown centre. DG18, 25 C: Sporulation moderate; texture fasiculate; mycelium white, conidia dull green; reverse yellow-orange. Micromorphology: Sclerotia absent. Conidiophores 100–200 lm long, with rough walled stipes, non- 103 vesiculate, terverticillate, branch up to 20 lm long, stipe 2.5–3.5 lm wide. Metulae 4–10, 9–15 9 2.5–4 lm. Phialides flask shaped, 3–8 per metula, 7.5–9.5 9 2–3 lm. Conidia in long, distorted chains, finely rough to rough, often with spiral striations, ovoidal to fusiform, often with a connective on one side, 2.5–3.5 lm. Colony diameters: 7 d, in mm: CYA 27–31; CYA15 C 33–37; CYA30 C 8–12; CYA37 C no growth; DG18 31–36; MEA 23–28; YES 34–39; CYAS 29–34; creatine agar 20–25, weak growth, poor acid production, base formation absent. Material examined: REPUBLIC OF KOREA, Deodeok root (Codonopsis lanceolata), Cheongyang, 2011, (CBS H-22504, holotype), ex-type living culture (CBS 140670 = DTO 327-C9 = EML-COD3 = IBT 33673). GenBank numbers: CAL: KP900994, ITS: KP941754, RPB2: MK033521, TUB2: KP900995. Notes: Penicillium psychrotrophicum belongs to series Corymbifera and is phylogenetically related to P. tulipae, P. radicicola and P. mali-pumilae, a new species described below. Penicillium psychrotrophicum differs from P. tulipae and P. radicicola (and P. hirsutum) by having smaller colonies on CYA, CYA incubated at 30 C and MEA. Penicillium mali-pumilae is phenotypically similar but differs by having smaller colonies on CYAS (29–34 vs 20–28 mm), DG18 (31–36 vs 21–31 mm) and CYA incubated at 15 C (33–37 vs 26–31 mm). Furthermore, P. psychrotrophicum produces (pale) yellow mycelium on YES, while P. mali-pumilae has white coloured mycelium and the reverse of P. psychrotrophicum on DG18 has a shade of orange and this is lacking in P. mali-pumilae. The taxonomy of P. tulipae and P. radicicola needs attention. Both species can be recognized using TUB2 sequences; however, these species can not be differentiated using CAL sequences only. Three groups of strains are present in the combined phylogram, though the relationship is not resolved. A further complication is the position of a strain (presumably) representing the type of P. virescens Bainier (non Sopp, 1912), CBS 109553 (= NRRL 974). This strain belongs to the P. tulipae-P. radicicola clade and the name might thus have priority over the other two. CBS 109553 differs from both P. tulipae and P. radicicola by having a (dark) brown reverse on CYA and CYAS and brown soluble pigment production. P. hirsutum, another member of series Corymbifera, also has a brown reverse on CYA; however, these reverses are in shades of yellow– brown or orange-brown. Sporulation on YES is poor and this feature is shared with P. radicicola, while P. tulipae colonies sporulate moderately or strong sporulation on YES. Base production on CREA is shared with P. tulipae and P. virescens; a feature absent in P. radicicola. More research is needed to resolve the taxonomy of this clade. 123 104 Fig. 68 Thysanorea uniseptata (MFLU 18-1723, holotype). a, b Colonies on natural substrate. c Conidiophore with conidia. d, e Conidiogenous cells with conidia. f Conidia. g Geminated conidium. 123 Fungal Diversity (2019) 96:1–242 h, i Colony on PDA (h from above view, i from below view). Scale bars: a = 250 lm, b = 100 lm, c = 50 lm, d, e = 10 lm, f, g = 5 lm Fungal Diversity (2019) 96:1–242 105 Fig. 69 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU and SSU sequence data representing Herpotrichiellaceae. Related sequences are taken from Liu et al. (2015), Tian et al. (2016) and Dong et al. (2018). Twenty-nine strains are included in the combined analyses which comprise 3634 characters (674 characters for ITS, 1361 characters for LSU, 1599 characters for SSU) after alignment. Cyphellophora oxyspora (CBS 698.73) and C. sessilis (CBS 243.85) (Cyphellophoraceae, Chaetothyriales) are used as the outgroup taxa. Single gene analyses were carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of - 11580.668860 is presented. The matrix had 664 distinct alignment patterns, with 28.94% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.255728, C = 0.221018, G = 0.274376, T = 0.248878; substitution rates AC = 1.635663, AG = 2.578277, AT = 1.796079, CG = 1.025171, CT = 6.468570, GT = 1.000000; gamma distribution shape parameter a = 0.513924. Maximum parsimony analysis of 3080 constant characters and 182 informative characters resulted in one equally most parsimonious tree (CI = 0.610, RI = 0.675, RC = 0.411, HI = 0.390). Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 75% are placed above and below the branches respectively. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are in bold. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Penicillium wandoense Hyang B. Lee, P.M. Kirk & T.T.T. Nguyen, sp. nov. Index Fungorum number: IF555411; Facesoffungi number: FoF05791; Fig. 78 Etymology: wandoense, referring to its collection location, Wan island (Wando) from where the species was isolated (Korea). Holotype: CNUFC-WT31-1 Macromorphology: CYA, 25 C: Sporulation on CYA moderate; colonies radially sulcate; surface texture velutinous; mycelium white; soluble pigment absent; reverse yellow brown to reddish brown; MEA, 25 C: Sporulation on MEA strong; colonies pale green; margins low, wide, irregular; velvety to floccose texture; reverse pale red or pale yellow; YES, 25 C: Sporulation poor, mycelium 123 106 Fungal Diversity (2019) 96:1–242 Fig. 70 Phylogram generated from maximum likelihood analysis based on combined TUB2, CAL, and RPB2 sequence data for Penicillium circulare and related species within the sect. Sclerotiora. Penicillium levitum (CBS 345.48) was used as outgroup taxon. Bootstrap values for maximum likelihood (ML) greater than 50% are placed above the branches. Type strains are in bold. The newly generated sequences are indicated in blue Fig. 71 Phylogram generated from maximum likelihood analysis based on combined TUB2, CAL, and RPB2 sequence data for Penicillium geumsanense and related species within the sect. Robsamsonia. Penicillium turbatum (CBS 237.60) was used as outgroup taxon. Bootstrap values for maximum likelihood (ML) greater than 50% are placed above the branches. Type strains are in bold. The newly generated sequences are indicated in blue 123 Fungal Diversity (2019) 96:1–242 107 Fig. 72 Phylogram generated from maximum likelihood analysis based on combined TUB2, CAL, and RPB2 sequence data for Penicillium psychrotrophicum and P. mali-pumilae and related species within the sect. Fasciculata series Corymbifera. Penicillium chrysogenum (CBS 306.48) was used as outgroup taxon. Bootstrap values for maximum likelihood (ML) equal to or greater than 70% are placed above the branches. Branches with Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are in bold. Type strains are in bold. The newly generated sequences are indicated in blue white; surface texture velutinous; reverse yellow brown to reddish brown. Micromorphology: Sclerotia absent. Conidiophores arising from agar surface, predominantly biverticillate, but also many monoverticillate, branch up to 26.5 lm long, stipe 2.5–3.5 lm wide. Metulae 2–4, 9.5–15.5 9 2.5–4.0 lm. Phialides ampulliform shaped, 2–5 per metula, 6.5–9.5 9 2–3.5 lm. Conidia subglobose to ellipsoidal, 3–4 9 2.5–3.5 lm. Culture characteristics: The isolate grew over a wide range of temperatures with varying growth rates on MEA, CYA, and YES. The average growth rates of CNUFCWT31-1 on MEA, CYA, and YES medium at 25 C were 34.5, 23, and 21 mm per 7 days, respectively. Optimal growth was observed around 25 C, slow growth was observed at below 10 C, and at 37 C. Material examined: REPUBLIC OF KOREA, Wando island (34190 1.2000 N 126450 0.0000 E), from freshwater, 24 August 2018 (CNUFC-WT31-1, holotype), ex-type living 123 108 Fungal Diversity (2019) 96:1–242 Fig. 73 Phylogram generated from maximum likelihood analysis based on combined TUB2, CAL, and RPB2 sequence data for Penicillium wandoense and related species within the sect. LanataDivaricata. Penicillium glabrum (CBS 125543) was used as outgroup taxon. Bootstrap values for maximum likelihood (ML) greater than 50% are placed above the branches. Type strains are in bold. The newly generated sequences are indicated in blue culture preserved as glycerol stock at - 80 C in the CNUFC. GenBank numbers: CAL: MK080564, MK080565, RPB2: MK080566, MK080567, TUB2: MK080562, MK080563. Notes: Penicillium wandoense differs from P. araracuarense, and P. wotroi by its pale red or pale yellow reverse on MEA. In the phylogenetic tree based on multiple genes, the strain formed a separate branch from other species of the genus and is considered to represent a new species. the Laboulbeniales, while the two new orders include oly a few of the known species (less than 3%). Laboulbeniomycetes Engl. Laboulbeniales Lindau Notes: The order Laboulbeniales includes about 2100 described species of obligate ectosymbionts on Arthropods. The Laboulbeniales sensu Lindau comprehended all the Laboulbeniomycetes, but recently the class has been splitted in three orders as a result of molecular analysis (Goldmann and Weir 2018; Haelewaters et al. 2019). However, the great majority of species is still classified in 123 Laboulbeniaceae G. Winter Notes: A full taxonomic treatment of the family is more than 30 years old (Tavares 1985), but this classification has been recently questioned in the papers reported above. Although a new comprehensive taxonomy the group has not yet been advanced, the two genera considered below can still be safely included in the family Laboulbeniaceae. Prolixandromyces R. K. Benjamin 1970, Aliso 7: 174 [MB 4369] = Monandromyces R. K. Benjamin 1999, Aliso 18: 72 [MB 28387] syn. nov. Notes: The genus Monandromyces was erected by R.K. Benjamin in 1999 to accommodate Autophagomyces microveliae Thaxt. and ten new species. The given diagnosis (page 72) is as follows: ‘‘Receptacle consisting of three in-line superposed cells (I, II, III) bearing distally on one side a stalked perithecium and on the other side a simple, free appendage; basal (I) and suprabasal (II) cells Fungal Diversity (2019) 96:1–242 109 Fig. 74 Penicillium circulare (CNUFC-GEU220-1, holotype). a, d Colonies in Czapek yeast agar (CYA). b, e Colonies in malt extract agar (MEA). c, f Colonies in yeast extract sucrose agar (YES). g–m Conidiophores. n Conidia. Scale bars: g–m = 20 lm, n = 10 lm strongly obliquely superposed; cell II subtending the perithecium; terminal cell (III) subtending the appendage. Appendage consisting of three superposed cells and a single, terminal, simple, flask-shaped antheridium; basal cell small, broader than long to nearly isodiametric; median cell elongate, several times longer than broad; terminal cell small, nearly isodiametric or ± elongate, always less than length of median cell; antheridium flask shaped with an 123 110 Fungal Diversity (2019) 96:1–242 Fig. 75 Penicillium geumsanense (CNUFC-GEU2229-1, holotype). a, d Colonies in Czapek yeast agar (CYA). b, e Colonies in malt extract agar (MEA). c, f Colonies in yeast extract sucrose agar (YES). g–m Conidiophores. n Conidia. Scale bars: g–m = 20 lm, n = 10 lm elongate, slender efferent tube. Perithecium with two stalk cells (VI, VII), three persistent basal cells (m, n, n’), and four vertical rows of outer wall cells of five cells each. ….’’. 123 The recent description of four new species in the genus Prolixandromyces (Kaishian and Weir 2018) and the consequent emendation of the genus, now including also a species with a single, terminal antheridium (i.e. Fungal Diversity (2019) 96:1–242 111 Fig. 76 Penicillium mali-pumilae (EML-MP6080-1, ex-type culture). a 7 d old cultures, 25 C, left to right; first row, all obverse, Czapek yeast agar (CYA), yeast extract sucrose agar (YES), dichloran 18% glycerol agar (DG18), malt extract agar (MEA); second row, CYA reverse, YES reverse, DG18 reverse, CREA obverse. b–f Conidiophores. g Conidia. Scale bars: b–g = 10 lm P. bromelicola A. Weir & Kaishian), removes any difference between Monandromyces, (characterized by a single antheridium) and Prolixandromyces. In the emended diagnosis (page 223) is written: ‘‘The uppermost cell in the series either producing a single, simple, terminal antheridium in the 3-celled appendage or more frequently 123 112 Fungal Diversity (2019) 96:1–242 Fig. 77 Penicillium psychrotrophicum (EML-COD3, ex-type culture). a 7 d old cultures, 25 C, left to right; first row, all obverse, Czapek yeast agar (CYA), yeast extract sucrose agar (YES), dichloran 18% glycerol agar (DG18), malt extract agar (MEA); second row, CYA reverse, YES reverse, DG18 reverse, CREA obverse. b–f Conidiophores. g Conidia. Scale bars: b–g = 10 lm in the 4–11-celled appendage giving rise to both a terminal and a lateral antheridium’’. All the other features are identical, as it is identical the family of most of the host insects: the Veliidae (semiaquatic bugs). As a consequence 123 Fungal Diversity (2019) 96:1–242 113 Fig. 78 Penicillium wandoense (CNUFC-WT31-1, holotype). a, d Colonies in Czapek yeast agar (CYA). b, e Colonies in malt extract agar (MEA). c, f Colonies in yeast extract sucrose agar (YES). g–m Conidiophores. n Conidia. Scale bars: g–m = 20 lm, n = 10 lm of the new synonymy all the species of Monandromyces are transferred to the genus Prolixandromyces with the following new combinations. Prolixandromyces australis (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces australis R.K. Benjamin 1999, Aliso 18: 78 [MB 459855]; Index Fungorum number: IF556540 123 114 Prolixandromyces elongatus (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces elongatus R.K. Benjamin 1999, Aliso 18: 86 [MB 459861]; Index Fungorum number: IF556541 Prolixandromyces falcatus (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces falcatus R.K. Benjamin 1999, Aliso 18: 77 [MB 459854]; Index Fungorum number: IF556542 Prolixandromyces longispinae (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces longispinae R.K. Benjamin 1999, Aliso 18: 87 [MB 459862]; Index Fungorum number: IF556543 Prolixandromyces microveliae (Thaxter) W. Rossi, comb. nov. : Autophagomyces microveliae Thaxter 1931, Mem. Amer. Acad. Arts Sci. 16: 96 = Monandromyces microveliae (Thaxter) R.K. Benjamin 1999, Aliso 18: 74 [MB 459852]; Index Fungorum number: IF556544 Prolixandromyces neoalardi (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces neoalardi R.K. Benjamin 1999, Aliso 18: 82 [MB 459859]; Index Fungorum number: IF556545 Prolixandromyces polhemorum (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces polhemorum R.K. Benjamin 1999, Aliso 18: 79 [MB 459856]; Index Fungorum number: IF556546 Prolixandromyces protuberans (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces protuberans R.K. Benjamin 1999, Aliso 18: 81 [MB 459858]; Index Fungorum number: IF556547 Prolixandromyces pseudoveliae (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces pseudoveliae R.K. Benjamin 1999, Aliso 18: 84 [MB 459860]; Index Fungorum number: IF556548 Prolixandromyces tenuistipitis (R.K. Benjamin) W. Rossi, comb. nov. : Monandromyces tenuistipitis R.K. Benjamin 1999, Aliso 18: 80 [MB 459857]; Index Fungorum number: IF556549 Prolixandromyces umbonatus (R.K. Benjamin) W. Rossi, comb. nov. 123 Fungal Diversity (2019) 96:1–242 : Monandromyces umbonatus R.K. Benjamin 1999, Aliso 18: 74 [MB 459853]; Index Fungorum number: IF556550 Stigmatomyces H. Karst., Chemismus Pfl.-Zelle: 78 (1869) Notes: After the recent description of Stigmatomyces coronatus (Rossi and Bernardi 2018), the accepted species are 148 distributed in all continents, all occurring on flies (Diptera) with the only exceptions being S. caudicola Speg. and S. euconni F. Picard, two species that certainly do not belong to the genus Stigmatomyces. Stigmatomyces succini is the only known fossil species of the Laboulbeniales reported so far (Rossi et al. 2005). Very few sequences are available for species in the genus. Stigmatomyces chamaemyiae W. Rossi & M. Leonardi, sp. nov. Index Fungorum number: IF828160; Facesoffungi number: FoF05175; Fig. 79 Etymology: From the name of the genus of the host insects, Chamaemyia. Holotype: FI WR3847 Receptacle, distal portion of the free appendage ad perithecial apex almost hyaline; stalk and basal cell of the free appendage and perithecial venter rusty or amber yellow; the rest of the fungus is very diluted reddish or amber yellow. Receptacle very variable in length, slightly tapering to the foot, with the suprabasal cell longer than the basal. Stalk cell of the appendage relatively small, distally prominent, strongly tapered below, its outer margin almost straight. Free appendage usually turned sidewise, with the axis consisting of five cells (rarely four), each bearing two anteridia, the uppermost of which is spinose. Stalk cell of the perithecium distinctly broader than long, separated by the secondary stalk-cell by a concave septum; the cells above it subequal in size and shape, the two outer of which are more or less prominent. Perithecial venter finely granular, oblong, slightly and symmetrically inflated, passing without abrupt transition to the spreading base of the paler and longer neck, which is slightly curved dorsally and slightly swollen at its junction with the tip; the latter is asymmetrical, with the ventral side more convex; the apex short, rather abruptly narrower, distinctly turned towards the dorsal side, ending in four, small, paired, subtruncate lips. Total length 205–440 lm; perithecium 45–75 9 130–220 lm; free appendage 50–65 lm; ascospores 30–35 lm. Material examined: MALTA, Ghajn Tuffieha Bay, sea level, 1 May 2001, B. Merz, on legs and apex abdomen of Chamaemyia flavicornis (Strobl) (Diptera, Chamaemyidae) (FI WR3533 & WR3534, paratypes). FRANCE, Pyrénées Orientales, La Bastide, Col de Palomère, high grassland, 10 June 2007, J. C. Deeming, on legs of C. juncorum (Fallén) (FI WR3142, paratype). PORTUGAL, Monte Gordo, V. Fungal Diversity (2019) 96:1–242 115 Fig. 79 Stigmatomyces chamaemyiae (FI WR3847, holotype). Scale bar = 50 lm R. S. Antonio, 11 May 2012, R. Andrade, on various parts of C. flavipalpis (Haliday) (FI WR3848, paratype); Sintra, Almargem do Bispo, 21 August 2013, R. Andrade, on various parts of C. polystigma (Meigen) (FI WR3847, holotype). Notes: The thalli on the tergites of C. flavipalpis from Portugal are very long (up to 590 lm) because of a very long receptacle (up to 350 lm). Stigmatomyces chamaemyiae is the first species of the Laboulbeniales reported on Chamaemyidae. Among the described species, the species most similar to S. chamaemyiae seems to be S. aphaniosomae W. Rossi & A. Weir occurring on Chyromyidae. In the latter species, however, the perithecial apex is very different, the free appendage is shorter, with the axis 123 116 Fungal Diversity (2019) 96:1–242 consisting of three cells, and the basal cell of the receptacle is usually longer than the suprabasal (Rossi and Weir 2011). Twenty-four species of Stigmatomyces are recorded from Portugal, while eleven are from France; to date, none of the species of Stigmatomyces have been reported from Malta (Castaldo et al. 2004; Rossi et al. 2013). Stigmatomyces cocksii W. Rossi & M. Leonardi, sp. nov. Index Fungorum number: IF828161; Facesoffungi number: FoF05176; Fig. 80 Etymology: Named after the Australian entomologist Graeme Cocks, who collected the new species. Holotype: CAMB WR4074a Thallus grayish yellow: the perithecial venter, the stalk cell and lower cell of the appendage are darker, the basal cell of the receptacle (cell I) and the free appendage are paler, while the suprabasal cell of the receptacle (cell II) is hyaline or almost so. Receptacle very variable in length; basal cell regularly tapering to the foot and rather abruptly constricted at the septum with the suprabasal, which is of nearly uniform diameter although variably elongate. Stalk cell of the appendage stout, broadly heart-shaped, distally prominent. Free appendage reaching the summit of the perithecial venter, with the axis consisting of four cells, the lower of which is about twice longer than broad, the others slightly broader than long and gradually smaller; each axial cell bears two antheridia except the uppermost, which bears a third, terminal antheridium. Perithecial stalk and basal cells relatively small, not strongly unequal in size, distinctly prominent externally. Perithecial venter regularly elliptical in outline, its surface distinctly warty, the wartlike elevations disposed transversely forming darker ridges; the neck abruptly narrower, longer than the venter, almost isodiametric, transversely striate, slightly enlarged at its junction with the tip, which tapers gradually to a subtruncate apex. Total length 175–345 lm; perithecial venter 38–62 9 40–58 lm; perithecial neck, up to155 lm; free appendage 48–57 lm; ascospores 29–31 lm. Material examined: AUSTRALIA, Queensland, Hermit Park, 44 Marks St., Malaise trap, 23 December 2015, G. Cocks, on Notiochyromya sp. nova (Diptera, Chyromyidae) (CAMB WR4074a, holotype; FI WR4074a, isotype), Same data as the type, 20 November 2015 (FI WR4073, paratype). Notes: Because of the transverse ridges formed by wartlike elevations, Stigmatomyces cocksii bears a superficial resembles to S. micrandrus Thaxt. and S. rugosus Thaxt., both parasitic on shore flies (Ephydridae). However, in the two latter species these ridges are smaller and more numerous, the appendage is distinctly shorter than the perithecial venter, the apex of perithecium bearing relatively large with diverging lips (Thaxter 1908). 123 Fig. 80 Stigmatomyces cocksii (CAMB WR4074a, holotype). Scale bar = 50 lm The only previously described species of Stigmatomyces on flies of the family Chyromyidae is S. aphaniosomae. The latter is very different from S. cocksii: among other distinguishing features, it has a smooth surface of the perithecial venter or almost so, a shorter perithecial neck, a shorter free appendage, the perithecial apex bearing Fungal Diversity (2019) 96:1–242 unequal and diverging lips (Rossi and Weir 2011). The identified species of Stigmatomyces reported from Australia thus far are three: S. limnophorae Thaxt., S. rugosus Thaxt. (Rossi et al. 2013), and S. zaleae W. Rossi & A. Weir (Rossi and Weir 2007). Stigmatomyces papei W. Rossi & M. Leonardi, sp. nov. Index Fungorum number: IF828162; Facesoffungi number: FoF05178; Fig. 81 Etymology: Named after the Danish dipterologist Thomas Pape, who collected the new species. Holotype: FI WR4044 Perithecium pale reddish brown; the rest of the thallus is almost hyaline or faintly tinged with reddish brown, with the exception of the lower cell of the free appendage which stands out with brilliant rusty red. Receptacle rather short and stocky, tapering throughout, the basal cell subtriangular with its content contracted above, the suprabasal distinctly longer. Stalk-cell of the appendage relatively long and narrow, distally prominent. Free appendage slender and attenuate, straight or sigmoid, its axis usually 117 consisting of 9 cells, each bearing a pair of slightly diverging antheridia arranged in two rows, except for the uppermost cell, which bears three superimposed antheridia. Stalk-cell of the perithecium large, longer than broad, its outer margin overlapped for most of its length by the secondary stalk-cell, which is only slightly smaller. Perithecial basal cells subequal, large, longer than broad. Perithecium longer than the rest of the thallus, slender and elongate, the venter oblong, the neck hardly distinguished, about as long as the venter, very slightly tapering towards the tip and ending with two large and unequal outgrowths, of which the shorter is broadly ovoid, the other beanshaped and bending outwards. Total length 575–640 lm; perithecium 60–70 9 340–380 lm; free appendage 105–140 lm; ascospores 55 lm. Material examined: TANZANIA, Morogoro Region, Udzungwa Mts. Natl. Park, Mizimu Camp, 769 m, 30–31 July 2015, T. Pape, N. Scharff & al., on tergites of Plagiostenopterina submetallica (Loew) (Diptera, Platystomatidae) (FI WR4044, holotype). Fig. 81 Stigmatomyces papei (FI WR4044, holotype). a Whole thallus. b Upper portion of a syntype. Scale bars: a, b = 50 lm 123 118 Fungal Diversity (2019) 96:1–242 Notes: A few thalli occurring on the upper surface of the right wing of the host are smaller and stockier, with the larger of the perithecial outgrowths longer and slenderer. The two large perithecial outgrowths are unique in the genus and make Stigmatomyces papei distinguishable at first sight. It is likely allied with S. ortalidanus Thaxt. described 100 years ago on Plagiostenopterina vicaria Hendel (= P. nigripes Enderlein) from the Cameroons and never reported since, with which it shares the structure of the free appendage, the shape of the perithecial venter and the elongation of the perithecial basal cells. However, the latter fungus lacks the two peculiar perithecial outgrowths, it is much longer (up to 1 mm), with very long suprabasal cell and perithecial stalk-cell, wall cells of the perithecium spirally twisted, and surface of the perithecial neck distinctly corrugated (Thaxter 1931). Stigmatomyces papei is the first species of Stigmatomyces reported thus far from Tanzania. Stigmatomyces tschirnhausii W. Rossi & M. Leonardi, sp. nov. Index Fungorum number: IF828163; Facesoffungi number: FoF05177; Fig. 82 Etymology: Named after the German dipterologist Michael von Tschirnhaus. Holotype: FI WR3542 Receptacle, distal portion of the free appendage ad perithecial apex almost hyaline; the rest of the fungus is unequally colored with diluted rusty red, with the perithecial venter and lower cell of the free appendage distinctly darker. Receptacle tapering throughout, the basal cell abruptly narrow just above the foot, the suprabasal longer, the horizontal septum dividing the two cells associated with a slight constriction. Stalk-cell of the appendage rather short and stout, broadly heart-shaped, very prominent distally. Axis of the appendage consisting of four cells, the lower distinctly larger and darker, about twice as long as broad, the others gradually smaller, each bearing a pair of relatively large antheridia with strongly curved necks, the uppermost of which is spinose. Stalk-cell of the perithecium flattened and subtriangular; the cells above distinctly prominent externally. Perithecial venter relatively short and stout, symmetrically inflated, the spiral wall cells separated by thin furrows; the neck distinctly longer, almost isodiametric above the slightly spreading base; the tip very slightly bent and tapering; the lips strongly unequal, two being short, flattened and divergent, the other two elongate, forming together a tooth-like, suberect prolongation. Total length 280–400 lm; perithecium 50–70 9 175–225 lm; free appendage 55–70 lm. Material examined: KENYA, Western Region, Kakamega District, Kakamega Forest, Kisere Forest, about 20 km NE’ Kakamega, about 90 km SSE’ Mount Elgon 123 Fig. 82 Stigmatomyces tschirnhausii (FI WR3542, holotype). Scale bar = 50 lm summit, 1597 m asl, forest, Malaise trap, 00230 0700 N, 34530 1600 E, 16 April–16 May 2003, M. Kraemer, on sternites near apex abdomen, antennae and legs of Dasiops sp. (Diptera, Lonchaeidae) (FI WR3542, holotype). Notes: The perithecial apex of Stigmatomyces tschirnhausii bearing two pairs of lips of very unequal length is reminiscent of the apex of S. dacinus Thaxt. The latter, occurring on Tephritidae, shares with the new species also the spirally twisted wall-cells of the perithecial venter, but is quite different in other features, such as the very long Fungal Diversity (2019) 96:1–242 stalk-cell of the appendage, the free appendage longer with the axis consisting of a higher number of cells, the basal of which is sterile, the twist of the perithecial venter extending to the neck (Thaxter 1931). Stigmatomyces tschirnhausii is the first species of this genus reported on flies of the family Lonchaeidae. The species of Stigmatomyces reported from Kenya to date are four: S. rugosus (Rossi 119 1988), S. limnophorae (Rossi et al. 2013), S. limosinae Thaxt. and S. dacinus (Rossi and Leonardi 2018). Stigmatomyces vikhrevii W. Rossi & M. Leonardi, sp. nov. Index Fungorum number: IF828164; Facesoffungi number: FoF05179; Fig. 83 Fig. 83 Stigmatomyces vikhrevii (FI WR3694, paratype). Scale bars = 50 lm 123 120 Etymology: Named after the Russian dipterologist Nikita Vikhrev, who collected the new species. Holotype: FI WR3691 Receptacle almost hyaline; the rest of the fungus is unequally colored with diluted rusty red, with the perithecial venter and lower cell of the free appendage distinctly darker. Receptacle rather stout, nearly uniform, with the basal cell abruptly tapered at the base and the suprabasal longer, sometimes slightly broader above. Stalkcell of the appendage short and stout, its outer margin more or less convex. Free appendage also short and stout, with the axis composed of three cells, the lower distinctly darker, each producing a pair of large antheridia, the uppermost of which bears a tiny spine. Stalk-cell of the perithecium much compressed, the secondary stalk-cell larger, with lower and outer convex margins. Perithecial venter elliptical in outline, its wall cells very slightly spiral with a verruculose surface, the verrucosity disposed more or less transversely; passing without abrupt transition to the base of the paler, smooth and shorter neck; the latter erect, its margins very slightly concave, distally broader at its junction with the stout, tapering tip; the subtruncate apex ends in four lips, three of which are very small, subequal and flattened, the fourth rounded and more prominent. Total length 210–415 lm; perithecium 45–60 9 140–220 lm; free appendage 50–55 lm; ascospores 40–45 lm. Material examined: MALAYSIA, Borneo, Sabah, Kota Kinabalu, 5.99 N, 116.09E, 26–30 December 2011, N. Vikhrev, on the tergites near the apex of the abdomen of Discomyza maculipennis (Wiedemann) (Diptera, Ephydridae) (FI WR3691, holotype; FI WR3692, WR3693, WR3694, paratype). Notes: Stigmatomyces vikhrevii bears a superficial resembles to S. compressus Thaxt. reported on Jamaican shore-flies, which however has a more slender habit, a different perithecial apex, and a much longer free appendage (Thaxter 1931). The species of Stigmatomyces reported thus far from Malaysia are seven. These are S. ilytheae Thaxt. (Thaxter 1917), S. dacinus Thaxt., S. limosinoides Thaxt., S. tortimasculus Thaxt. (Thaxter 1918), S. sygmoideus Thaxt., S. subinflatus Thaxt. (Thaxter 1931) and S. neurochaetae W. Rossi et A. Weir (Rossi and Weir 2007). Actually, S. subinflatus was described twice: in the first description the host (Drosophila sp.) is reported from Cameroon (Thaxter 1918, p. 740), while in the second the collecting place was transferred to ‘‘Sarawak, Borneo’’ (Thaxter 1931, p. 134). The right place seems the second because this is the one reported on the label of the type slide of S. subinflatus (Thaxter no. 2180 = FH 6409). 123 Fungal Diversity (2019) 96:1–242 Lecanoromycetes O.E. Erikss. & Winka Notes: This class represents most of the lichenized members within Ascomycota (Lücking et al. 2017). However, it also includes some non-lichenized ascomycetes (Kirk et al. 2008). Most of the taxa are characterised by apothecial ascomata, while a few form perithecia (Ekanayaka et al. 2017). Subclass Lecanoromycetidae Notes: The subclass Lecanoromycetidae with in Lecanoromycetes was established by Hibbett et al. (2007). This subclass comprises Caliciales, Lecanorales, Lecideales, Leprocaulales, Peltigerales, Rhizocarpales and Teloschistales (Jaklitsch et al. 2016). Caliciales Bessey Notes: This order includes both mazaediate and nonmazaediate genera of both crustose, fruticose and foliose lichenized genera in various habitats (Jaklitsch et al. 2016; Ekanayaka et al. 2017). However the taxa in this order are highly varied in their morphological characters. Caliciaceae Chevall. Notes: This family was established by Chevallier (1826) and currently comprises 29 genera and 630 species (Jaklitsch et al. 2016; Wijayawardene et al. 2018a). Taxa are mostly lichenized and form crustose to squamulose, foliose, or fruticose lichen thalli. Ascomata are mostly apothecial (Jaklitsch et al. 2016). Species are widely distributed in temperate, subtropical, and tropical regions, especially on bark, rocks and wood (Jaklitsch et al. 2016). Buellia De Not. Notes: The lichen genus Buellia was introduced by De Notaris (1846). Currently this genus includes around 400 species. The genus is characterized by black lecideine apothecia, septate, oblong to ellipsoid, rarely citriform, hyaline to brownish ascospores and a deep reddish brown to yellow or yellowish green to rarely hyaline hypothecium (Joshi et al. 2010). Buellia viridula Ekanayaka & K.D. Hyde, sp. nov. Index Fungorum number: IF555467; Facesoffungi number: FoF05082; Fig. 84 Etymology: The specific epithet viridula refers to the pigment colour in hymenium layer of apothecia. Holotype: MFLU 16-0587 Saprobic on dead stems. Sexual morph Apothecia 300– 350 9 110–160 lm ( x = 331.7 9 132.6 lm, n = 10), arising singly, erumpent, sessile, black when fresh. Receptacle pulvinate, black. Disc flat to slightly convex, black. Margins black. Excipulum 22–38 lm ( x = 31.5 lm, n = 10) wide at margins and flanks, composed of black cells of textura angularis. Hymenium blackish green. Paraphyses 1–3 lm wide ( x = 2.1 lm, n = 20), numerous, branched, septate. Fungal Diversity (2019) 96:1–242 121 Fig. 84 Buellia viridula (MFLU 16-0587, holotype). a Substrate. b, c Ascomata on wood. d Cross section of an ascoma. e Close up of the cross section of ascoma at margins. f Cylindrical paraphyses. g–j Cylindric-clavate asci. k–n Obvoid ascospores. Scale bars: b = 500 lm, d = 150 lm, c = 100 lm, e = 60 lm, f = 20 lm, g– j = 10 lm, k–n = 5 lm Asci 24–50 9 8–14 lm ( x = 34.3 9 10.4 lm, n = 30), 8spored, globose, long stipitate, apex rounded, inoperculate, amyloid ring absent at the apex. Ascospores 7–11 9 3–4 lm ( x = 9.5 9 3.7 lm, n = 40), multiseriate, obovoid, immature spore hyaline, greenish brown at maturity, smooth-walled, 1-septate, guttules present. Asexual morph undetermined. Material examined: THAILAND, Chiang Rai Province, Doi Mae Suai, on dead stems, 22 June 2015, A.H. Ekanayaka, HD0026 (MFLU 16-0587 holotype; HKAS 104247, isotype). GenBank numbers: ITS: MK075947 (MFLU 16-0587a), MK075948 (MFLU 16-0587-b). Notes: Buellia viridula is characterized by sessile black apothecia, a blackish excipulum, a blackish green 123 122 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 85 Phylogram generated from a maximum likelihood analysis based on ITS sequence data. The newly generated nucleotide sequences are compared against the GenBank (http://www.ncbi.nlm. nih.gov/) database using the Mega BLAST program. Related sequences were obtained from GenBank. Fifty-eight strains were included in the sequence analyses, which comprised 655 characters including gaps. Calicium abietinum (MP40) is used as the outgroup taxon. The best scoring RAxML tree with a final likelihood value of - 9417.010515 is presented. The matrix had 460 distinct alignment patterns, with 21.50% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.223, C = 0.287, G = 0.258, T = 0.232; substitution rates AC = 1.600208, AG = 2.778296, AT = 1.882342, CG = 0.827224, CT = 6.521407, GT = 1.000000; gamma distribution shape parameter a = 0.407728. Bootstrap support values for ML equal or greater than 50% are given above the nodes. The newly generated sequences are in blue hymenium, globose asci and ellipsoid greenish ascospores. Buellia viridula is similar to B. ocellata and B. jugorum by having coloured hymenium, but the latter two species have ornamented ascospores (Scheidegger 1993). The phylogenetic analysis clearly establishes B. viridula as a new species (Fig. 85). Lecanorales Nannf. Notes: This order was established by Nannfeldt (1932). Taxa form lichen thalli with protococcoid green photobionts (Crespo et al. 2010; Miadlikowska et al. 2014). The members are widely distributed (Kirk et al. 2008; Ekanayaka et al. 2017). Lecanoraceae Körb. Notes: This family was established by Körber (1855). The members of the family are widely distributed and commonly found on rocks, soil or bark (Kalb et al. 2011). Species are characterised by apothecial ascomata, sparingly branched paraphyses, semifissitunicate cylindrical to clavate asci, and ellipsoid, subglobose or bacilliform, hyaline ascospores (Kalb et al. 2011). Lecidella yunnanensis Ekanayaka, & K.D. Hyde, sp. nov. Index Fungorum number: IF555468; Facesoffungi number: FoF05083; Fig. 86 Etymology: The specific epithet yunnanensis refers to the province where type specimen collected. Holotype: MFLU 17-1910 Saprobic on dead stems. Sexual morph 300– 500 9 150–200 lm ( x = 368 9 181 lm, n = 10). Apothecia arising singly or in small groups, sessile, slightly erumpent from the substrate, pulvinate, black with shiny appearance. Receptacle convex. Margin distinct, black. Hymenium upper part dark brown and lower part hyaline, within a thick gelatinous matrix. Ectal excipulum composed of small, thin-walled, blackish cells of textura angularis, with algal cells. Medullary excipulum composed of narrow, long, thin-walled, hyaline cells of textura 123 intricata. Paraphyses 1–2 lm wide ( x = 1.6 lm, n = 20), numerous, filiform, propoloid, septate, slightly swollen at the apices, apices are glued together to develop a pseudoepithecium. Asci 40–50 9 9–13 lm ( x = 45 9 12 lm, n = 30), 8-spored, cylindric–clavate, narrowed to the base, short pedicellate, rounded at the apex, amyloid ring absent at the ascus apex, croziers present at the base of asci. Ascospores 10–15 9 5–6 lm ( x = 12 9 5.4 lm, n = 40), hyaline when immature, greenish brown at maturity, smooth walled, ellipsoid, 1-septate, guttulate. Material examined: CHINA, Kunming, Yunnan Province, 25 April 2016, A.H. Ekanayaka, HC 001, HC004 (MFLU 17-1910, holotype). GenBank numbers: ITS: MK075945, LSU: MK075949 (MFLU 17-1910-a); ITS MK075946, LSU MK075950 (MFLU 17-1910-b). Notes: Lecidella yunnanensis is characterized by black, sessile, pulvinate apothecia, with a black disk with a shiny appearance, propoloid septate paraphyses, cylindric–clavate asci and ellipsoid ascospores. Lecidella yunnanensis is similar to Lecidella enteroleucella, however L. enteroleucella differs from L. yunnanensis by having black, greenish to reddish brown to olive disks and a hyaline hymenium layer (Nash et al. 2004). In the phylogenetic analysis, L. yunnanensis groups clearly distinct from known species of Lecidella (Fig. 87). Pilocarpaceae Zahlbr. Notes: Pilocarpaceae is a family of mainly crustose lichens, including mostly foliicolous lineages, plus the bulk of species currently classified in the genus Micarea, previously placed in a separate family Micareaceae (Andersen and Ekman 2004, 2005; Miadlikowska et al. 2014). Micarea Fr. Notes: The genus Micarea itself is highly heterogeneous, with several lineages falling outside the genus in its proper sense and even outside the Pilocarpaceae; the monophyletic core group also possibly represents more than one genus (Czarnota and Guzow-Krzemińska 2010; Schmull et al. 2011; Ekman and Svensson 2014). Here, we introduce a new species with squamulose thallus that forms an isolated lineage within the monophyletic Micarea core clade (Fig. 89). It might well represent a separate genus, but further data are required to resolve this entire clade taxonomically. Micarea squamulosa Aptroot, Lücking & M. Cáceres, sp. nov. Index Fungorum number: IF555463; Facesoffungi number: FoF05966; Fig. 88 Etymology: Referring to the squamulose thallus. Holotype: M. E. S. Cáceres & A. Aptroot ISE 40719 (ISE) 123 124 Fungal Diversity (2019) 96:1–242 Fig. 86 Lecidella yunnanensis (MFLU 17-1910, holotype). a Substrate. b Ascomata on wood. c Cross section of an ascoma. d Close up of a vertical section of the ascoma at margin. e Aseptate paraphyses. f–i Short pedicellate asci. j–m Ovate ascospores. Scale bars: c, d = 100 lm, e = 20 lm, f–i = 15 lm, j–m = 6 lm Terricolous squamulose Micarea with loose to agglutinated, olive-green lobes which are flat to terete and solid to hollow, without rhizines; apothecia immarginate, dark brown, ascospores hyaline, ellipsoid, 11–12 9 6–6.5 lm. Thallus squamulose, terricolous, extending up to 2 cm diam., consisting of loose to crowded and occasionally agglutinated, simple to plicate or lobate flat to ascending squamules of c. 0.4–1.2 mm diam. Squamules entire to shallowly incised, partly saddle-shaped to erect, olive green, not pruinose, smooth but not shiny; margins rather thick and rounded, of the same colour as the upper surface; lower surface dull, pale olive green, without rhizines; medulla whitish, often becoming hollow, up to 300 lm thick; marginal lobes more often flat, solid and elongate than the central lobes which are more often erect and hollow. Photobiont green. Ascomata apothecia, superficial on the thallus, solitary and centrally on the squamules, round to irregular in outline, 0.2–0.6 mm diam.; disc dark brown to dark grey, convex, dull; margin c. 0.1 mm wide, dark brown (paler inside), glossy, not visible from above. Hymenium hyaline, c. 60 lm high; epihymenium olive brown; excipulum dark brown outside, hyaline inside; hypothecium contiguous with the excipulum, hyaline. Paraphyses rather agglutinated. Asci cylindrical, with 8 ascospores. Ascospores hyaline, ellipsoid, 11–12 9 6–6.5 lm, thin-walled, without gelatinous sheath 123 Fungal Diversity (2019) 96:1–242 Fig. 87 Phylogram generated from a maximum likelihood analysis based on combined ITS and LSU sequence data. The newly generated nucleotide sequences are compared against the GenBank database using the Mega BLAST program. Related sequences were obtained from GenBank. Fifty-eight strains were included in the sequence analyses, which comprised 1397 (ITS-1-607, LSU-608-1397) characters including gaps. Rhizoplaca porterii 55145 is used as the outgroup taxon. The best scoring RAxML tree with a final likelihood 125 value of - 5739.681991 is presented. The matrix had 462 distinct alignment patterns, with 31.86% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.234, C = 0.258, G = 0.284, T = 0.224; substitution rates AC = 1.101172, AG = 2.447560, AT = 1.586081, CG = 0.813290, CT = 8.452097, GT = 1.000000; gamma distribution shape parameter a = 0.205506. Bootstrap support values for ML equal or greater than 50% are given above the nodes. The newly generated sequences are in blue Fig. 88 Micarea squamulosa (isotype). Habitus. Scale bar = 0.5 mm 123 126 Fungal Diversity (2019) 96:1–242 but usually with a central vacuole/oil droplet. Pycnidia not observed. Secondary chemistry: Cortex and medulla K–, P–, C–, UV–. TLC: Atranorin, gyrophoric acid and skyrin; all apparently in concentrations which are too low for spot reactions. Material examined: BRAZIL. BAHIA: Chapada Diamantina, Lençóis, Cachoeira do Mosquito; 12230 S, 41220 4000 W, alt. c. 450–500 m; 22 July 2017, on siliceous soil, M.E.S. Cáceres & A. Aptroot ISE 40719 (ISE holotype; ABL isotype). Notes: This species looks strongly like a small species of Toninia, especially because of the olive-green, often contorted squamules with rounded margins and often hollow medulla. Timdal (1991) did not accept any of the ten species originally described from Brazil in this genus; they were partly referred to other genera by him, and partly excluded without alternative generic dispositions. Aptroot (2002) referred some of these to the genera Bacidiopsora and Micarea. The first genuine Toninia, T. massata (Tuck) Herre, was reported from Brazil only recently by Cáceres et al. (2017), from Ceará state. The present species differs from all previously reported toninioid species from the country. Sequence data of the mitochondrial small subunit rDNA (GenBank accession: MK080110) place this taxon in the Pilocarpaceae, in the monophyletic Micarea core clade (Fig. 89); however, no distinctly squamulose taxa were known in this clade before. This clade remains unresolved and may include several genera, apart from Micarea s.str., which is represented by the strongly supported M. prasina subclade. In that case, the squamulose taxon may represent a separate genus. The species previously reported as Toninia from Brazil or similar taxa elsewhere also need to be resampled, to test whether some of these are related to the M. bullata lineage. The species was found in the Chapada Diamantina, an area in the state of Bahia in tropical northeastern Brazil representing the second largest conserved area in the country outside the Amazon (Funch et al. 2009). The natural vegetation of the Chapada Diamantina includes various types of evergreen moist and semi-deciduous dry tropical forest, belonging to the Caatinga, Cerrado and Atlantic rain forest biomes and the Restinga-like Capitinga (Funch et al. 2009). The siliceous bedrock, geologically mostly of precambrian origin, surfaces in many parts in the form of isolated inselbergs surrounded by forest. These inselbergs are usually only up to 500 m higher than the surrounding landscape, but the meso- and microclimate on the often rather flat tops and the generally very steep slopes Fig. 89 Best-scoring maximum-likelihood tree of a subset of Pilocarpaceae based on mtSSU sequence data, including the Pilocarpaceae core (as outgroup or sister clade) and the Micarea core clade, showing the position of the new species, M. squamulosa. GenBank accession numbers are indicated in the tree figure. The final alignment comprised 914 columns, with 450 distinct alignment patterns, and the final likelihood was –6458.814273. Estimated base frequencies were as follows: A = 0.335039, C = 0.150256, G = 0.201024, T = 0.313681, and substitution rates as follows: AC = 1.022403, AG = 4.972129, AT = 1.612594, CG = 0.507417, CT = 4.968323, GT = 1.000000, with the gamma distribution shape parameter a = 0.290696. Bootstrap values based on 1000 pseudoreplicates are placed above the branches 123 Fungal Diversity (2019) 96:1–242 differs markedly from that in the surrounding areas, leading to the presence of so-called campos rupestres, to some extent resembling elfin forest but dominated by sclerophyllous small trees and shrubs adapted to the particular edaphic and climatic conditions of this habitat (Conceição et al. 2007; Funch et al. 2009). So far, around 300 lichenized species have been reported from the region (Kalb 1986; Ahti and Oksanen 1990; Kashiwadani and Kalb 1993; Mendonça et al. 2016; Aptroot and Cáceres 2018a, b; Kalb and Aptroot 2018), many from a single field trip during which also the following enigmatic squamulose lichen was collected, on vertical soil at a raised ledge in the middle of a river. The new species was found on wet sand with bryophytes and a possibly undescribed Gyalideopsis on a vertical ledge, in crevices between sandstone bedrock, on a small island in a broad river near a waterfall. The location is very wet and poor in nutrients. Ostropomycetidae V. Reeb, Lutzoni & Cl. Roux Notes: Ostropomycetidae is the second largest subclass in class Lecanoromycetes, including six orders with mostly lichenized taxa but also some secondarily delichenized lineages (Lücking et al. 2017). Ostropales Nannf. Notes: Ostropales is the largest order of subclass Ostropomycetidae, with 11 families, approximately 140 genera, and well over 3200 accepted species (Lücking et al. 2017). Among the largest families are Graphidaceae and Porinaceae, with a new genus in Graphidaceae and a new species in Porinaceae introduced below. Porinaceae Rchb. Notes: Porinaceae is a mid-sized family of approximately 360 known species, currently classified in six genera (Lücking et al. 2017). The generic concept is in flux and based on molecular data, at least 15 genus-level lineages can be distinguished (Sobreira et al. 2018). All Porinaceae produce perithecioid ascomata and are rather uniform in their internal morphology; some lineages also reproduce vegetatively by means of isidia or isidioid structures and rarely soredia. The latter are characteristic of the genus Myeloconis (McCarthy and Elix 1996) but otherwise unknown in the family. Here we describe a new species with exposed, black perithecia and a sorediate thallus (Fig. 90) which based on ITS sequence data belongs in the Porina byssophila clade (Fig. 91), representing an unnamed genus (Sobreira et al. 2018) corresponding to Pseudosagedia subgen. Limosagedia Hafellner & Kalb (1995). Porina sorediata Aptroot, Lücking & M. Cáceres, sp. nov. Index Fungorum number: IF555464; Facesoffungi number: FoF05967; Fig. 90 127 Etymology: Referring to the sorediate thallus, an unusual feature in the family except in the genus Myeloconis. Holotype: M. E. S. Cáceres & A. Aptroot ISE 42451 (ISE) Porina with capitate soralia and black semiglobose perithecia with 3-septate, somewhat curved ascospores with rounded ends, of 14–15 9 3.5–4 lm. Thallus covering araes of bark of up to 20 cm diam., mostly smooth and often somewhat glossy, greyish green, not surrounded by a differentiated prothallus; photobiont tentepohlioid, globiose to usually ellipsoid, c. 7–12 9 6–8 lm. Soralia numerous, round, globose, somehwat paler than the thallus, especially when abraded, c. 0.2–0.5 mm diam. Soredia farinose, globose to usually ellipsoid, c. 17–27 9 10–18 lm, consisting of one trentepohlioid algal cell of c. 7–12 9 6–8 lm, surrounded by dozens of hyaline, semiglobose to somewhat angular hyphal cells with lumina of c. 2–3 lm. Ascomata uncommon, perithecia, black, semiglobose, c. 0.2 mm diam., c. 0.1 mm high. Wall clearly differentiated into a proper ascoma wall and an involucrellum which covers the exposed half of the ascoma with a separate layer; both layers purple-brown, unchanged in KOH. Ascospores 8/ascus, hyaline, 3-septate, somewhat curved with rounded ends, 14–15 9 3.5–4 lm. Material examined: Brazil, Alagoas, Quebrangulo, Reserva Biólogica de Pedra Talhada, private area, 9150 S, 36250 3500 W, on tree bark, c. 600 m alt., 21–23 Oct 2017, M.E.S. Cáceres & A. Aptroot ISE 42451 (ISE holotype; ABL isotype); Same locality and datails, ISE 42328, 42450 & 42675 (ISE, ABL). Notes: Only three species in the family Porinaceae are known to form soredia, all in the genus Myeloconis (McCarthy and Elix 1996). Species of the latter genus have brighly pigmented soredia and medulla, perithecia with pale walls, and large, muriform ascospores, thus quite different from the new species. Species with exposed, black perithecia occur in several lineages of Porinaceae, including the genera Pseudosagedia and Trichothelium and the Porina byssophila clade (Sobreira et al. 2018). Unfortunately, we were unable to generate mtSSU data for the present taxon, and the nuLSU primers sequenced a fungal contaminant. However, the ITS data (GenBank accession: MK080109) place the new species in the P. byssophila clade, together with P. guentheri (Fig. 91). This is notable, as the P. byssophila clade, in spite of its very different morphology, is closely related to the genus Myeloconis based on mtSSU data (Sobreira et al. 2018), and hence the molecular phylogeny brings the lineages with soredia rather close together. Hafellner and Kalb (1995) established subgenus Limosagedia for a subgroup of species within Pseudosagedia, with Porina linearis (Leight.) Zahlbr. as type but including also P. byssophila. The stated difference was the K ? bluish outer perithecial wall. Thus, 123 128 Fungal Diversity (2019) 96:1–242 Fig. 90 Porina sorediata (isotype). a Thallus with ascomata and soralia. b Thallus with soralia. c Section through ascoma. d Soredia. Scales bars: a, b = 1 mm, c = 0.2 mm, d = 50 lm Fig. 91 Best-scoring maximum-likelihood tree of available subset of Porinaceae based on ITS sequence data, showing placement of P. sorediata in the P. byssophila clade which, based on mtSSU sequence data, is closely related to the genus Myeloconis (Sobreira et al. 2018). The final alignment comprised 534 columns, with 210 distinct alignment patterns, and the final likelihood was –1903.538679. Estimated base frequencies were as follows: A = 0.187911, C = 0.338406, G = 0.295727, T = 0.177956, and substitution rates as follows: AC = 1.312730, AG = 3.551014, AT = 2.797834, CG = 1.299677, CT = 6.511396, GT = 1.000000, with the gamma distribution shape parameter a = 0.330010. Bootstrap values based on 1000 pseudoreplicates are placed above the branches Limosagedia could potentially be a name available for this clade when raised to genus level; however, in the present species no such K-reaction was observed. The new taxon was found in the Pedra Talhada Biological Reserve in the state of Alagoas, northeastern Brazil, where it was locally quite common on shaded bark of various trees, but it has 123 Fungal Diversity (2019) 96:1–242 not been observed in any other site in this locality, a phenomenon also known from other species that appear to represent local endemics. Pedra Talhada Biological Reserve is ecologically quite unique, representing one of the larger remnants of Atlantic rain forest in northeastern Brazil, with various phytophysiognomies formed on inselbergs built from precambrian gneiss (Nusbaumer et al. 2015; Studer et al. 2015). The new species is thus far only known only from Brazil. Graphidaceae Dumort. Notes: Graphidaceae is the second largest family of lichenized Ascomycota, containing containing over 2000 known species and about 80 genera (Lücking et al. 2017). Although most lineages have been sequenced, odd taxa that do not fit any of the genera now recognized continue to be encountered on a regular basis (Lumbsch et al. 2014). Here we describe a new genus based on sequenced material found in the same area as Porina sorediata, described elsewhere in this paper. The species was identified as the rare Thelotrema cryptotrema Nyl., later recombined as Ocellularia cryptotrema (Nyl.) Kalb and more recently as Schizotrema cryptotrema (Nyl.) Rivas Plata & Mangold (Rivas Plata et al. 2010). Sequence data show that the taxon is unrelated to Thelotrema or Schizotrema and instead forms a novel lineage in tribe Ocellulariae, related to the Ocellularia profunda and O. praestans clades (Fig. 93). The latter two are not members of Ocellularia sensu stricto and will require new genera placement as well, but here we formally describe a new genus for Thelotrema cryptotrema, which in addition to its phylogenetic placement also exhibits some unusual features within tribe Ocellularieae (Fig. 92). Cryptoschizotrema Aptroot, Lücking & M. Cáceres, gen. nov. Index Fungorum number: IF555465; Facesoffungi number: FoF05968 Etymology: Referring to the anatomical resemblence with Schizotrema species, due to the partially carbonized and concentrically layered, fissured excipulum, and the notion that the Schizotrema-like excipulum is hidden beneath a covering layer. Type species: Cryptoschizotrema cryptotrema (Nyl.) Aptroot, Lücking & M. Cáceres. Cryptoschizotrema cryptotrema (Nyl.) Aptroot, Lücking & M. Cáceres, comb. nov. Index Fungorum number: IF555466; Facesoffungi number: FoF05969; Fig. 92 = Thelotrema cryptotrema Nyl., Ann Sci Nat Bot Sér 5, 7: 318 (1867). = T. secoligella Müll. Arg., Hedwigia 34: 31 (1895). 129 = T. annulatum Müll. Arg., J Linn Soc Bot 30: 453 (1895). New genus in Graphidaceae with myriotremoid to porinoid, immersed-erumpent ascomata with a narrow pore, dark brown to partially (apically) carbonized, fissured and becoming concentrically layered excipulum lacking periphysoids, muriform, subdistoseptate, I-negative (nonamyloid) ascospores with thin walls and septa, and psoromic acid chemistry. Thallus light greenish grey, unevenverrucose, with dense, prosoplectenchymatous cortex. Ascomata immersed-erumpent, rounded, disc covered by narrow pore, invisible; proper margin covered by whitish thalline layer, not visible, when abraded seen as brown to black, fissured lobules forming concentric layers in older ascomata. Columella absent. Excipulum prosoplectenchymatous, brown to partially (apically) carbonized, becoming layered (striate) in older ascomata; periphysoids absent. Hymenium clear; paraphyses unbranched. Ascospores 1–2(–4)/ascus, richly muriform, up to 100 9 30 lm, oblong-ellipsoid, distoseptae with rather thin walls and septa and angular lumina, colorless, I– (non-amyloid). Secondary chemistry: psoromic acid (thallus P ? yellow). Material examined: Brazil, Alagoas, Quebrangulo, Reserva Biólogica de Pedra Talhada, private area, 9150 S, 36250 3500 W, on wooden pole, c. 550 m alt., 21–23 Oct 2017, M.E.S. Cáceres & A. Aptroot ISE 42655 (ABL, B, ISE). Notes: Tribe Ocellularieae largely comprises the taxa previously classified in the collective genera Myriotrema and Thelotrema (Hale 1980; Lumbsch et al. 2014). Both are highly polyphyletic, as shown by Frisch et al. (2006) and Rivas Plata et al. (2012), and even after formally segregating several lineages as new genera, quite a number of orphaned lineages remain unnamed. Our results show that Thelotrema cryptotrema (nuLSU Genbank accession: MK080108) represents an additional lineage in this heterogeneous assembly, related to the O. profunda and O. praestans clades. This topology would theoretically allow to include the latter two clades together with T. cryptotrema in a single genus, for which the here established name would then be available. However, at this time we refrain from placing additional species in Cryptoschizotrema, since the long branches leading to each clade (Fig. 93) do not suggest a single genus to be involved and the differences between Cryptoschizotrema and the other two clades are substantial, including the massive, fully carbonized and non-striate excipulum and the distoseptate, I ? violet-blue (amyloid) ascospores in O. profunda and O. praestans and relatives. This monospecific new genus is thus far only known from Central and South America. It is tropical lowland to montane species growing on bark or wood; the sequenced specimen was found in Brazilian Atlantic Rain Forest, on a 123 130 Fungal Diversity (2019) 96:1–242 Fig. 92 Cryptoschizotrema cryptotrema. a–d Cáceres & A. Aptroot ISE 42655. e, f isotype of T.annulatum in BM). a Thallus with ascomata. b Ascomata enlarged. c Hymenium with asci and ascospores. d Ascospores. e Abraded ascoma showing schizotremoid proper margin (photograph A. Mangold). f Section through ascoma showing striate, apically carbonized excipulum (photograph A. Mangold). Scale bars: a = 1 mm, b, e = 0.5 mm, c, d = 10 lm, f = 100 lm wooden pole of a fence dividing different parts of the reserve. Wijayawardene et al. (2018a) accepted 11 orders and 44 families in Leotiomycetes. Leotiomycetes O.E. Erikss. & Winka Notes: Leotiomycetes is the largest the largest class of inoperculate discomycetes, a group of nonlichenized ascomycetes (Eriksson and Winka 1997; Lumbsch et al. 2005; Zhang and Wang 2015; Ekanayaka et al. 2017). Helotiales Nannf. Notes: Helotiales species are abundant in humid areas as saprobes on dead leaves and shoot on herbaceous and woody plants (Webster 2007). In addition, they occur as plant pathogens and endophytes. Helotiales is a highly 123 Fungal Diversity (2019) 96:1–242 131 Fig. 93 Best-scoring maximum-likelihood tree of subset of Graphidaceae tribe Ocellularieae based on three markers (mtSSU, nuLSU, RPB2), showing placement of Cryptoschizotrema cryptotrema near the O. profunda and O. praestans clades. The final alignment comprised 2618 columns, with 983 distinct alignment patterns, and the final likelihood was –14732.747014. Estimated base frequencies were as follows: A = 0.289260, C = 0.202949, G = 0.260731, T = 0.247061, and substitution rates as follows: AC = 0.929088, AG = 3.966253, AT = 2.046683, CG = 0.790669, CT = 7.857784, GT = 1.000000, with the gamma distribution shape parameter a = 0.231216. Bootstrap values based on 1000 pseudoreplicates are placed above the branches diversified order and accommodates 27 families (Ekanayaka et al. 2017; Wijayawardene et al. 2018a). Saprobic on dead flower petals and leaves of Rosa sp. Sexual morph Undetermined. Asexual morph Conidiomata sporodochial, 175–200 lm high ( x = 185 lm, n = 10), 122– 140 lm diam. ( x = 133.5 lm, n = 10), 48–56.5 lm ( x = 52.5 lm, n = 5) wide at the base, superficial, scattered, sessile, relatively sphaerical, initially yellowish brown, turning brownish or almost black, initially convex cap turning to concave. Peridium thick, textura angularis. Conidiophores 11.2–19.8 9 1.0–1.4 lm ( x = 15.6 9 1.2 lm, n = 10), a dense palisade, hyaline to pale brown, tapering distally. Conidiogenous cells 2.2–3.3 9 0.8–1.1 lm ( x = 2.6 9 0.9 lm, n = 10), enteroblastic, phialidic, hyaline, tapering towards the apex, terminal and lateral. Conidia 4.8– 7.4 9 1.5–2.4 lm ( x = 6.1 9 1.9 lm, n = 40), L/W 3.2, 1-celled, hyaline, fusoid to allantoid, slightly falcate, apex acute, base obtuse, thin-walled, guttules at both ends, formed singly on the conidiogenous cells. Culture characteristics: Colonies reaching to 36.5 mm after 2 weeks at 28 C on PDA with two zones. Inner zone dirty yellow to light brown, flat, slightly clock wise rotate sparse mycelia, reverse light brown; outer layer dirty white, flat, smooth, lobate margin, reverse dirty white. Later, zonate, dirty white, flat surface, smooth with light brown patches, reverse dark brown centre and light brown margin. Chaetomellaceae Baral, P.R. Johnst. & Rossman Notes: Chaetomellaceae was introduced by Baral et al. (2015) with its type genus Chaetomella Fuckel to accommodate non-lichanized discomycetes as a family insertae sedis in Leotiomycetes. ITS and LSU based phylogenetic study revealed that the family consists of four genera viz. Chaetomella (= Zoellneria), Pilidium Kunze (= Discohainesia), Sphaerographium Sacc. and Synchaetomella Decock & Seifert. Wijayawardene et al. (2018a) accepted the family in the Helotiales (Leotiomycetes). Pilidium Kunze Notes: The genus Pilidium was introduced with P. acerinum (Alb. & Schwein.) Kunze as the type species with a Hainesia synasexual, and Discohainesia sexual morphs (Rossman et al. 2004; Johnston et al. 2014). Currently, 22 accepted Pilidium species have been reported as saprobes and pathogens on various hosts (Wijayawardene et al. 2017a). Pilidium concavum (Desm.) Höhn., Sber. Akad. Wiss. Wien, Math.-naturw. Kl., Abt. 1 124: 148 (1915) Facesoffungi number: FoF05212; Fig. 94 123 132 Fungal Diversity (2019) 96:1–242 Fig. 94 Pilidium concavum (MFLU 18-0106, new geographical record). a Host. b, c Appearance of conidiomata on host surface. d Vertical section through the conidioma. e–h Conidiophores and conidiogenous cells. i, j Culture on PDA (i from above view, j from below view). k Conidiospores. Scale bars: b–d = 100 lm, e–h, k = 5 lm Material examined: THAILAND, Chiang Rai Province, Muang District, Tha Sut Sub-district, (20 030 24.700 N, 99 520 23.500 E), dead petals and leaves of Rosa sp. (Rosaceae), 20 August 2017, MC. Samarakoon, SAMC017 (MFLU 18-0106, HKAS 102341), living culture (MFLUCC 17-2671). GenBank numbers: ITS: MK100327, LSU: MK108192, SSU: MK108189. Notes: Pilidium concavum (synanamorph Hainesia lythri (Desm.) Höhn.; sexual morph Discohainesia oenotherae (Cooke and Ellis) Nannf.), is an interesting species with morphological differentiation due to genetic control (Sutton 1980; Palm 1991; Rossman et al. 2004). The species is often pathogenic and occasionally saprobic on flowers, fruits and oil and forest plants (Cardin et al. 2009; Geng et al. 2012; Lopes et al. 2010; Ayoubi et al. 2016). Pilidium concavum has been recorded mainly in temperate regions and on many hosts. Palm (1991) reports P. concavum on Rosa leaves from the United Kingdom and Connecticut (USA). The strain isolated from petals and leaves of Rosa sp. in this study is similar in morphology and phylogeny to P. concavum. The ITS-LSU multi-locus phylogenetic studies reveal that strain MFLUCC 17-2671 is closely related (86% ML/0.99 BYPP) to P. concavum and the P. lythri clade (Fig. 95). Morphological characterizations, especially the conidia, overlap with the consistency measurements (5.5–7.2 9 1.4–1.8 lm) described by Palm (1991). Thus, we report this as P. concavum on petals and leaves of Rosa sp. as a saprobe from Thailand, a new geographical record. 123 Pezizomycetes O.E. Erikss. & Winka Notes: Pezizomycetes is the representative operculate discomycete of Ascomycota. It is characterized by apothecial ascomata or deformed apothecia and asci with an apical slit or operculum (Hansen and Pfister 2006). Presently, 21 families are accepted in this class (Ekanayaka et al. 2018). Phylogenetic studies based on multiple genes Fungal Diversity (2019) 96:1–242 133 Fig. 95 Phylogram generated from maximum likelihood analysis based on combined ITS and LSU sequence data for Pilidium species. Related sequences are taken from Marin-Felix et al. (2017). Thirteen strains are included in the combined analyses which comprise 1250 characters (465 characters from ITS, 785 characters for LSU) after alignment. Chaetomella raphigera (BPI 843551) (Chaetomellaceae, Helotiales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood value of - 3008.149666 is presented. The matrix had 128 distinct alignment patterns, with 1.3% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.258502, C = 0.211927, G = 0.278401, T = 0.251171; substitution rates AC = 3.685184, AG = 2.311128, AT = 2.325507, CG = 0.504055, CT = 5.692118, GT = 1.000000; gamma distribution shape parameter a = 0.02. Bootstrap values for maximum likelihood (ML) equal to or greater than 70% and Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are placed respectively on the branch. Type strains are bold. The newly generated sequence is indicated in blue indicate the Pezizomycetes is a monophyletic group (Ekanayaka et al. 2018). (Korf & W.Y. Zhuang) Y.J. Yao & Spooner, B. luteola Svrček, B. nicholsonii (Massee) Spooner & Y.J. Yao, B. ovalispora (Boud.) Van Vooren, B. subprolata (Korf & W.Y. Zhuang) Y.J. Yao & Spooner and B. vermiphila Brumm. & R. Kristiansen are known. Three species have available molecular data (Perry et al. 2007; Hansen et al. 2013; Kušan et al. 2018). Pezizales J. Schröt. Notes: Phylogenetic analyses by Ekanayaka et al. (2018) show six clades in this order. All operculate taxa and some non-apothecial taxa are included in Pezizales. In addition to genera from 21 families with a resolved placement, 15 genera are unresolved and listed as Pezizales incertae sedis (Jaklitsch et al. 2016; Wijayawardene et al. 2017a). Ascodesmidaceae J. Schröt. [as ‘Ascodesmidacei’] Notes: The family Ascodesmidaceae was introduced by Schröter (1893) and is typified by Ascodesmis Tiegh. with Ascodesmis aurea Tiegh. as type species. This family is characterized by small ascomata, straight or curved paraphyses and smooth or ornamented ascospores (Güngör et al. 2014; Ekanayaka et al. 2018). Currently, twelve genera are listed in this family, but just four have sequence data. (Ekanayaka et al. 2018; Wijayawardene et al. 2018a). Phylogenetically, this family is close to Pyronemataceae Corda and Otideaceae Eckblad within Pezizales (Ekanayaka et al. 2018). Boubovia Svrček Notes: The genus Boubovia was established by Svrček (1791) for Boubovia luteola Svrček which is characterized by glabrous to setaceous ascomata, inamyloid asci and smooth or ornamented ascospores with a cyanophilic sheath (Korf and Zhuang 1984; Yao and Spooner 1996; Kristiansen 2016). Six species, Boubovia ascoboloides Boubovia gelatinosa M. Zeng, Q. Zhao & K.D. Hyde, sp. nov. Index Fungorum number: IF555488; Facesoffungi number: FoF05174; Fig. 96 Etymology: The epithet refers to ascospores surrounded by gelatinous sheath. Holotype: HKAS 102407 Saprobic on sheep dung. Sexual morph Ascomata superficial, scattered. Disc pulvinate, yellowish to orange, shiny, up to 1 mm high, 0.5 mm broad, with 1–2 mm long setaceous hairs, exceeding the disc. Medullary excipulum of textura intricata. Ectal excipulum, hyaline, J-, comprised of 8–12 9 6–8 lm cells of textura angularis to globulosa. Paraphyses 2–3 lm broad, filiform, septate, straight, hyaline, J-. Asci 108–160 9 18–23 lm, 8-spored, subcylindrical to clavate, operculate, inamyloid. Ascospores [20/1/1, in H2O] (16.3–)16.6–18.9(– 21.1) 9 (7.7–)8.9–10.4(–11.0) (Q = 1.65–2.30, Q = 1.84 ± 0.15), ellipsoid to subglobose, uniseriate to biseriate or irregular, smooth-walled, enveloped by gelatinous sheath or over 2/3 part surrounded by sheath, deBarybubbles present. Asexual morph Undetermined. 123 134 Fungal Diversity (2019) 96:1–242 Fig. 96 Boubovia gelatinosa (HKAS 102407, holotype). a–c Typical mature specimens. d Squash of ascomata in Melzer’s reagent. e Receptacle surface of pileus. f Asci and paraphyses in Melzer’s reagent. g–i Asci (h, i Asci in Melzer’s reagent). j–m Ascospores. Scale bars: d = 100 lm, f–i = 30 lm, e, j = 20 lm, k–m = 15 lm Material examined: CHINA, Jiuzhaigou, Sichuan, on sheep dung, 22 August 2018; Ming Zeng, ZM1d (HKAS 102407, holotype). GenBank numbers: ITS: MK093850, LSU: MK093851, SSU: MK093852, TEF1-a: MK093853. Notes: This species is distinguished by orange shiny ascomata with long hairs, straight paraphyses and ellipsoid, smooth wall ascospores with sheath. The phylogenetic tree shows a close relationship with Coprotus ochraceus (P. Crouan & H. Crouan) J. Moravec (Fig. 97), but morphologically, this species is more similar to Boubovia species. Boubovia may yet prove to be synonym of Coprotus or more than two genera may exist here, but further collections are needed. Boubovia vermiphila differs in its 123 Fungal Diversity (2019) 96:1–242 135 Fig. 97 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU, SSU and TEF1-a sequence data for Boubovia species and several closely related genera in Ascodesmidaceae. Related sequences are taken from Perry et al. (2007), Hansen et al. (2013), Lindemann et al. (2015) and Kušan et al. (2018). Sixteen strains are included in the combined analyses which comprise 3584 characters (669 characters for ITS, 890 characters for LSU, 1043 characters for SSU, 982 characters for TEF1-a) after alignment. Pseudoboubovia benkertii (V.K. 3202) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of - 12041.925019 is presented. The matrix had 891 distinct alignment patterns, with 51.43% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.246462, C = 0.237728, G = 0.271829, T = 0.243980; substitution rates AC = 0.807419, AG = 1.258958, AT = 1.238671, CG = 0.599522, CT = 3.934755, GT = 1.000000; gamma distribution shape parameter a = 0.201448. Maximum parsimony analysis of 2717 constant characters and 600 informative characters resulted in two equally most parsimonious tree of 1614 steps (CI = 0.778, RI = 0.704, RC = 0.547, HI = 0.222). Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 75 are placed above and below the branches respectively. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.90. The newly generated sequence is indicated in blue glabrous ascomata, curved paraphyses, ornamented ascospores with 1–2 oil drops and large apical ascospore caps (Brummelen and Kristiansen 1999; Kristiansen 2016). Boubovia ascoboloides and B. ovalispora differ in their glabrous ascomata and curved paraphyses (Korf and Zhuang 1984; Kristiansen 2016). Boubovia subprolata 123 136 differs in its pinkish glabrous ascomata and curved paraphyses (Korf and Zhuang 1991). Sordariomycetes O.E. Erikss. & Winka Diaporthomycetidae Senan., Maharachch. & K.D. Hyde Notes: The subclass Diaporthomycetidae was introduced to accommodate ten orders and 26 families (Maharachchikumbura et al. 2015). Diaporthales, initially placed in the subclass Sordariomycetidae, where it formed a sister group with Magnaporthaceae, has been transferred and established as type order of the subclass Diaporthomycetidae (Zhang et al. 2006). Species of this subclass are saprobic, endophytic, parasitic or pathogenic in plants, animals or humans (Maharachchikumbura et al. 2015). The sexual morphs comprise stromata made up of orange, brown or black parenchymatous tissues and seldom pyriform ascomata. The ascospores are 2–3 seriate, aseptate or septate, of different shapes, and may or may not have appendages or sheaths. The asexual morphs are ceolomycetes, where conidiomata are acervuli or pycnidial with annellidic or phialidic proliferating conidiogenous cells, or as hyphomycetes, where conidiophores originate from hyphae with occasionally coloured terminal or lateral conidiogenous cells (Maharachchikumbura et al. 2015). Diaporthales Nannf. Notes: Diaporthales is a well-supported monophyletic clade within the Diaporthomycetidae and members of this order are saprobes and soil inhabitants, pathogens, parasites or endophytes in plants (Zhang et al. 2006; Rossman et al. 2007; Maharachchikumbura et al. 2015). Some are also pathogenic to humans and animals (Rossman et al. 2007). Diaporthalean microfungi are currently accommodated within 21 families. Among all the species in this order, Cryphonectria parasitica (Murrill) M.E. Barr is probably the most well-known chestnut blight fungus which destroyed the whole population of American chestnut trees [Castanea dentata (Marsh.) Borkh.] within a short period of time (Anagnostakis 1987). Taxa belonging to this order have yellowish brown to black valsoid or diatrypoid pseudo- or asco-stromata, brown to black perithecial fruiting bodies, regularly immersed in stromata of leafy or woody substrates with papillate ostioles swollen at the apex. The mature 2–32 spored asci with a J- refractive apical apparatus, float freely and become dehiscent at maturity (Barr 1978; Samuels and Blackwell 2001; Senanayake et al. 2017). The asexual states are usually ceolomycetous, with conidiophores arising from the topmost cell layer of basal or parietal tissue or from under the developing scutellum; otherwise, they are reduced to enteroblastic, holoblastic, phialidic or annellidic hyaline or olivaceous conidiogenous cells (Sogonov et al. 2008; Senanayake et al. 2017). 123 Fungal Diversity (2019) 96:1–242 Diaporthaceae Höhn. Notes: The family Diaporthaceae currently comprises 14 genera (Maharachchikumbura et al. 2016; Senanayake et al. 2017, 2018; Dissanayake et al. 2017b). Diaporthe Nitschke Notes: Diaporthe includes economically important plant pathogens, endophytes, and saprobes (Udayanga et al. 2014; Hyde et al. 2016). They have a broad range of hosts from cultivated crops to ornamental plants (Dissanayake et al. 2017b). Rossman et al. (2015a) prioritized the older name Diaporthe over Phomopsis, while the latest account is provided in Fig. 100. An updated phylogenetic tree for Diaporthe species with new records and new species is presented in Fig. 100. Diaporthe italiana Chethana, Camporesi & K. D. Hyde, sp. nov. Index Fungorum number: IF555376; Facesoffungi number: FoF04934; Fig. 98 Etymology: The specific epithet ‘italiana’ was given after the country, where the fungus was collected. Holotype: MFLU 17-0311 Saprobic on dead aerial branch of Morus alba L. Sexual morph Perithecia 0.2–0.42 mm ( x = 0.311 mm, n = 10), solitary, scattered, immersed to semi-immersed, partially erumpent, globose, black, with tapering perithecial necks. Peridium multi-layered, 20 lm wide at the top, 10–12 lm wide in sides, outer layer composed of 4–5 layers of thick, brown cells, inner 3–4 layers of hyaline cells of textura angularis, cells towards inner layer lighter. Asci 45– 55 9 6–8.5 lm ( x = 51 9 7.7 lm, n = 20), 8-spored, unitunicate, sessile, cymbiform to clavate. Ascospores 10.5–13 9 2.7–4.5 lm ( x = 12.1 9 3.5 lm, n = 40), overlapping biseriate, hyaline, fusiform to allantoid, slightly curved, 2-celled, widest near the centre, with acute rounded ends, 1-median septate, rarely bi-guttulate, often tetra-guttulate, with larger guttules at centre and the smaller towards the ends, smooth-walled. Asexual morph not observed. Culture characteristics: Colonies on the PDA, circular, fluffy, aerial, white mycelium with fimbriate margin, olivaceous to grey olivaceous reverse, relatively slow growing, reach 3 cm diam. after 7 days at 25 C. Material examined: ITALY, Forlı̀-Cesena Province, Forlı̀, via Friuli, on dead, aerial branch of Morus alba L. (Moraceae), 19 January 2017, Erio Camporesi, IT 3217 (MFLU 17-0311, holotype), ex-type living culture (MFLUCC 18-0090, KUMCC 18-0002); (HKAS 101460, isotype), ex-isotype living culture (MFLUCC 18-0091, KUMCC 18-0003). GenBank numbers: CAL: MH853690, ITS: MH846237, TEF1-a: MH853686, TUB2: MH853688 (MFLUCC Fungal Diversity (2019) 96:1–242 137 Fig. 98 Diaporthe italiana (MFLUCC 18-0090, holotype). a Appearance of perithecia on dead branch of Morus alba. b Longitudinal section of a perithecium. c Longitudinal section of a perithecium wall showing cell organization. d Bi-seriate asci. e–j Ascospores. k, l Colony on PDA (k from above view, l from below view). Scale bars: a = 1 mm, b = 100 lm, c = 20 lm, d = 10 lm, e–j = 5 lm 18–0090); CAL: MH853691, ITS: MH846238, TEF1-a: MH853687, TUB2: MH853689 (MFLUCC 18-0091). Notes: Diaporthe italiana is phylogenetically closely related to D. rudis, but is clearly distinguished based on the morphology and phylogeny. In our phylogenetic analysis of combined ITS, TEF1-a, TUB2, and CAL sequence data of Diaporthe species (Fig. 100), our strains (18-0090 and 18-0091) cluster together with the ex-type strain of Diaporthe rudis (AR3422) with relatively high bootstrap and Bayesian probabilities (100% ML/1.00 BYPP). Sequence comparison for ITS region between D. rudis (AR3422) and D. italiana (MFLUCC 18–0091) showed 2.11% base pair differences. Our collection is distinct from D. rudis in having smaller asci (45–55 9 6–8.5 lm), and smaller ascospores (10.5–13 9 2.7–4.5 lm), in contrast to larger asci (53.5–58.5 9 10.6–12 lm), and larger ascospores (12–14.2 9 3.5–3.7 lm) of D. rudis (Udayanga et al. 2014). A Diaporthe species named Diaporthe mori Berl. (current name Phomopsis moricola (Sacc.) Grove) has been identified from Morus alba and M. nigra (BargagliPetrucci 1915). However, they differ from our species in having smaller, immersed perithecia (0.18–0.2 mm), smaller asci (36–40 9 7 lm), and smaller, bi-guttulate ascospores (7–8 9 3–3.5 lm). Since, this species were introduced in 1915, DNA sequences were unavailable in 123 138 the databases. Therefore, they were not included in the phylogenetic analysis. Diaporthe rumicicola Manawasinghe, Camporesi & K.D. Hyde, sp. nov. Index Fungorum number: IF555379; Facesoffungi number: FoF04940; Fig. 99 Etymology: Name reflects the host genus. Holotype: MFLU 18-0739 Saprobic on dead leaves of woody plants. Sexual morph Undetermined. Asexual morph Conidiomata on host pycnidial, 98–280 lm diam. ( x = 208 lm, n = 10), ampulliform, scattered, immersed, ostiolate, with elongate Fig. 99 Diaporthe rumicicola (MFLU 18-0739, holotype). a Appearance of conidiomata on dead branch of Rumex spp. b Cross section of conidioma on host. c Pycnidial wall on host. d, e Conidiogenous cells with developing conidia. f Conidia on host. g Pycnidia on PDA. 123 Fungal Diversity (2019) 96:1–242 black neck. Pycnidial wall comprising 2–3 layers, heavily pigmented, thick-walled, comprising dark brown cells of textura angularis, with lighter cells towards the inside, with inner layer composed of 2–5 layers, hyaline, thinwalled cells of textura angularis. Conidiophores on host reduced to conidiogenous cells. Pycnidia in culture black, immersed, solitary or merged, exudated. Paraphyses observed on culture. Conidiogenous cells enteroblastic with percurrent annellations, integrated, clustered, hyaline, smooth-walled. Alpha conidia, on host 3–5 9 2–3 lm ( x = 3.5 9 2.5 lm, n = 30), hyaline, smooth-walled, mono or bi-guttulate. Beta conidia not observed. h Appearance of paraphysis. i, j Colony on PDA (i from above view, j from below view). Scale bars: a = 2000 lm, b = 200 lm, g = 100 lm, c–e, h = 20 lm, f = 10 lm Fungal Diversity (2019) 96:1–242 Culture characteristics: Colonies on PDA 50 mm diam., after 7 days at 25 C, cream to white cottony mycelium, irregular and lobate at the margins, white at the centre; reverse yellowish and concentric lines on conidial formation can be observed when colony getting old. Material examined: ITALY, near Castrocaro Terme (province of Forlı́-Cesena), on dead aerial stem of Rumex sp. (Polygonaceae), 19 February 2018, Camporesi Erio IT (MFLU 18-0739, holotype), ex-type living culture (MFLUCC 18-1566, JZB320005). GenBank numbers: ITS: MH84623, TEF1-a: MK049554, TUB2: MK049555. Notes: Diaporthe rumicicola was isolated from dead branch of Rumex sp. from Forı́-Cesena of Italy. Morphologically, this species is similar to genus Diaporthe, and the combined gene sequence analysis of ITS, TEF1-a, CAL and TUB2 genes placed the taxon in Diaporthe (Fig. 100). The current species develop a sister clade with Diaporthe foeniculina with 100% bootstrap value in ML analysis and 1.0 Bayesian probabilities in Bayesian analysis (Fig. 99). Morphologically the current species is differing with developing smaller alpha conidia (3.5 9 2.4 lm) compared to D. foeniculina (8.5–9 9 2.3–2.5 lm) (Udayanga et al. 2014). In addition to that this is the first time that Diaporthe species recorded on Rumex sp. (Farr and Rossman 2019). Gnomoniaceae G. Winter Notes: The family Gnomoniaceae, typified by the genus Gnomonia Ces. & De Not., encompasses fungi with immersed black perithecia which are solitary or aggregated with or without a stromatic tissue (Sogonov et al. 2008). The neck of the perithecial bodies extends beyond the host epidermis and may be short or long with an upright, oblique or parallel positioning to the host surface (Sogonov et al. 2008). The asexual morphs of species of Gnomoniaceae produce acervular or pycnidial fruiting bodies (Monod 1983). Species belonging to this family have mainly been reported from north America and Europe, with only few originating from Asia (Walker et al. 2014). Most of the species are host-specific (Senanayake et al. 2017) and commonly occur as endophytes, pathogens or saprobes, even though, they sometimes behave as parasites or opportunistic pathogens (Sieber 2007; Senanayake et al. 2018). Plagiostoma Fuckel Notes: Plagiostoma (Gnomoniaceae, Diaporthales) encompasses microfungal members of pathogenic, endophytic or saprobic nature inhabiting stems, branches, twigs and leaves of woody and herbaceous plants belonging to various families which occur mostly temperate regions of the northern hemisphere as well as some parts of Asia (Mejı́a et al. 2011; Walker et al. 2014). The genus, typified 139 by P. euphorbiae (Fuckel) Fuckel, currently accommodates 32 species (Mejı́a et al. 2011; Walker et al. 2014; Senanayake et al. 2017). These species produce perithecial and pycnidial fruiting bodies on overwintered leaves and twigs (Walker et al. 2014) and the genus primarily distinguishes itself from other genera in the family by neck characters of the perithecia along with ascospore morphology (Mejı́a et al. 2011). Other than these, presence or absence of stroma, geographical location or host are secondary differentiating features (Mejı́a et al. 2011). In this study, we report Plagiostoma salicellum (Fr.) Sogonov for the first time from a new geographical area in Europe, notably Italy. Plagiostoma salicellum (Fr.) Sogonov, Stud. Mycol. 62: 73 (2008) Facesoffungi number: FoF05761; Fig. 101 Saprobic on dead branches of Salix caprea. Sexual morph Ascomata 205–350 9 260–330 lm ( x = 275 9 290.5 lm, n = 6), perithecial, mostly solitary, scattered, immersed in bark, producing a faint elevation of the periderm, each comprising of a convergent, protruding neck, black, subglobose. Neck 110–120 lm long ( x = 113 lm, n = 6), 76–126 lm diam. at the base ( x = 97.3 lm, n = 6) and 88–131 lm diam. at apex ( x = 103.8 lm, n = 6), cylindrical, eccentric to lateral, straight to very slightly twisted, not surrounded by a stroma. Peridium 18–29 lm wide, thick, comprising 4–6 layers, outer layer heavily pigmented, thick walled, consisting of dark brown to blackish cells of textura angularis, cells towards inside lighter, inner layer composed of 1–2 layers, pale brown to hyaline, flattened, thin-walled cells of textura angularis. Asci 32–55 9 5–15 lm ( x = 43.7 9 8.7 lm, n = 34), 8-spored, cylindrical to clavate, clavate-elongated, straight to slightly curved, broadly rounded apex with a conspicuous J- refractive apical ring, 1.7–4.3 lm diam., often consisting of a short but not necessarily persistent, tapered and pointed stipe. Ascospores 12–18 9 1–4 lm ( x = 15.8 9 2.5 lm, n = 54), obliquely parallel or irregularly seriate, hyaline, ellipsoidelongated, tapering slightly towards rounded ends, 1-septate, slightly constricted at median to sub-median septum, with granular cytoplasm, without any sheath or appendage. Asexual morph Undetermined. Material examined: ITALY, Province of Forlı̀-Cesena [FC], Monte Fumaiolo, on dead aerial branch of Salix caprea L. (Salicaceae), 27 August 2017, Erio Camporesi, IT 3454 (MFLU 17-1601). GenBank numbers: ITS: MK080111, TEF1-a: MK424970, TUB2: MK424971. Notes: Plagiostoma salicellum is characterised by a cylindrical neck surrounded by a white stroma and narrow ellipsoid-elongated ascospores often having short appendages (Mejı́a et al. 2011). However, in our specimen no 123 140 Fig. 100 Phylogenetic tree generated by maximum likelihood analysis of combined ITS, TEF1-a, TUB2 and CAL sequence data of Diaporthe species. Sequences were obtained from GenBank. Onehundred and ninety-two strains are included in the analyses, which comprise 2603 characters including gaps. Single gene analyses are carried out to compare the topology of the tree and clade stability. Tree was rooted with Diaporthella corylina (CBS 121124). Tree topology of the Bayesian analysis was similar to the RAxML. The best scoring RAxML tree with a final likelihood value of = 123 Fungal Diversity (2019) 96:1–242 - 60797.442221 is presented. The matrix had 2107 distinct alignment patterns, with 40.98%of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.228179, C = 0.282550, G = 0.268109, T = 0.221162; substitution rates AC = 0.854574, AG = 3.369047, AT = 0.975280, CG = 0.800183, CT = 3.251377, GT = 1.000000; gamma distribution shape parameter a = 0.881490. RAxML and Bayesian posterior probabilities values C 60% (ML) and 0.9 (BYPP) are shown respectively near the nodes. The newly generated sequences are indicated in blue Fungal Diversity (2019) 96:1–242 141 Fig. 100 continued stroma was observed around the perithecial neck which was also shorter x = 113 lm (this study) vs x = 177 lm (Mejı́a et al. 2011)] and broader both at the base x = 97.3 lm (this study) vs x = 81 lm (Mejı́a et al. 2011)] and at the apex x = 103.8 lm (this study) vs x = 79 lm (Mejı́a et al. 2011)] as compared to the lectotype and epitype of P. salicellum described by Mejı́a et al. (2011). The perithecia were mostly solitary rather than aggregated and narrower in diameter x = 290.5 lm (this study) vs x = 397 lm (Mejı́a et al. 2011)]. The asci also were shorter and narrower ( x = 43.7 9 8.7 lm (this study) vs x = 55.0 9 13.0 lm (Mejı́a et al. 2011)]. The ascospores 123 142 Fungal Diversity (2019) 96:1–242 Fig. 100 continued of P. salicellum (MFLU 17-1601) did not have an appendage unlike those of the lectotype and epitype of P. salicellum and they were smaller x = 15.8 9 2.5 lm (this study) vs x = 18.5 9 3.5 lm (Mejı́a et al. 2011)]. However, the length of the ascospores may not be a key distinguishing factor since Mejı́a et al. (2011) observed that 123 length and width of ascospores of P. salicellum may vary, even within an ascus. The combined data set of ITS, TUB2 and TEF1-a sequences in our phylogenetic analyses has not clearly separated our strain (MFLU 17-1601) with significant bootstrap support from the strains of P. salicellum (AR 3828) and (LCM 449.01) (Fig. 102). The morphological Fungal Diversity (2019) 96:1–242 Fig. 101 Plagiostoma salicellum (MFLU 17-1601, new geographical record). a Appearance of ascomata on host substrate. b Close-up of ascoma on host substrate. c Section of an ascoma. d Section through 143 ostiole. e Peridium. f–h Asci. i–l Ascospores. Scale bars: a = 500 lm, b, c = 200 lm, d = 100 lm, i–l = 20 lm, e–h = 10 lm 123 144 Fig. 102 Phylogram generated from maximum likelihood analysis based on combined ITS, TUB2 and TEF1-a sequence data for Plagiostoma species in Gnomoniaceae. Related sequences were taken from Sogonov et al. (2008), Mejı́a et al. (2011) and Walker et al. (2014). Sixty-five strains are included in the combined analyses which comprise 1948 characters (511 characters for ITS, 777 characters for TUB2, 660 characters for TEF1-a) after alignment. Apiognomonia errabunda (AR 4182) (Gnomoniaceae, Diaporthales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood value of - 12805.418798 is presented. The matrix had 982 distinct alignment patterns, with 17.73% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.220722, C = 0.299844, G = 0.226468, T = 0.252965; substitution rates AC = 0.958759, AG = 3.395596, AT = 0.951907, CG = 1.027587, CT = 4.491882, GT = 1.000000; gamma distribution shape parameter a = 0.409235. Bootstrap values for maximum likelihood (ML) equal to or greater than 60 are placed above and below the branches. Bayesian posterior probabilities (BYPP) equal to or greater than 0.90 are in bold. The newly generated sequence is indicated in bold and blue 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 differences mentioned above may be the result of phenotypic plasticity arising from a need to adapt to the climatic conditions in Italy (Jeewon & Hyde 2016). Due to lack of significant molecular support, we therefore report our strain as P. salicellum, collected for the first time from Italy. It is also known from Canada, Europe (Austria, Belgium, Bulgaria, Czech Republic, Germany, Poland, Sweden, Switzerland, UK and USA (Sogonov et al. 2008). 145 Distoseptispora K.D. Hyde, McKenzie & Maharachch. Notes: Distoseptispora is typified by D. fluminicola Mckenzie, H.Y. Su, Z.L. Luo & K.D. Hyde (Su et al. 2016) and currently 16 species are recorded in Index Fungorum (2019). In addition, D. adscendens (HKUCC 10820) and D. leonensis (HKUCC 10822) were also included in this genus based on evidence of phylogeny (Su et al. 2016). Members in this genus are commonly encountered in freshwater habitats and are hyphomycetous, which have olivaceous, brown or yellowish/reddish brown conidia, euseptate or distoseptate, obclavate or cylindrical with rounded apex, and are of indeterminate length (Su et al. 2016; Luo et al. 2018; Tibpromma et al. 2018; Yang et al. 2018b). macronematous, mononematous, unbranched, erect, straight or flexuous, cylindrical, smooth-walled, brown to pale olivaceous, septate, slightly constricted at septa, thinwalled. Conidiogenous cells 13–20 lm long ( x = 16 lm, n = 10), 3–5 lm wide ( x = 4.5 lm, n = 10), monoblastic or polyblastic, terminal, determinate, cylindrical, brown to pale olivaceous. Conidia 17–30 lm long ( x = 25 lm, n = 15), 7.5–10 lm wide ( x = 9.2 lm, n = 15), acrogenous, solitary, obpyriform to obclavate, broad cylindrical or irregular, straight or curved, 3–5-distoseptate, olivaceous, thick-walled, smooth, rounded at the apex, truncate at the base. Culture characteristics: On PDA, colony circular, reaching 40 mm in 45 days at 25 C, grey from above, dark grey from below, surface rough, dry, raised, edge entire. Material examined: CHINA, Dehong, on submerged wood in a stream, 25 November 2017, G.N. Wang H10A (HKAS 101738, holotype), ex-type living culture (KUMCC 18-0090). GenBank numbers: ITS: MK085061, LSU: MK079662, TEF1-a: MK087659. Notes: Distoseptispora dehongensis is phylogenetically, a strongly supported member in Distoseptisporaceae (100 ML/1.00 BYPP) in Fig. 105. Morphological characteristics of D. dehongensis (such as cylindrical, unbranched, septate conidiophores and distoseptate conidia) matches well with those in Distoseptispora. It can be easily distinguished from the other taxa in Distoseptispora by its fewer distosepta (3–5-distoseptate vs. [ 7-distoseptate) and shorter length (17–30 lm vs. [ 35 lm) of conidia (Hyde et al. 2016; Su et al. 2016; Yang et al. 2018b; Luo et al. 2018). Distoseptispora dehongensis has similar conidial dimensions to D. martinii. However, they differ in conidial shape (Xia et al. 2017). In our multi-gene phylogenetic analyses, D. dehongensis has a close phylogenetic affinity to D. obpyriformis and D. rostrata with weak support (Fig. 105), and they can be distinguished in conidial shape, size and number of distosepta (Luo et al. 2018). Distoseptispora dehongensis W. Dong, H. Zhang & K.D. Hyde, sp. nov. Index Fungorum number: IF555406; Facesoffungi number: FoF05057; Fig. 103 Etymology: Name reflects Dehong, from where the species was isolated. Holotype: HKAS 101738 Saprobic on decaying wood submerged in freshwater. Sexual morph Undetermined. Asexual morph Colonies gregarious or scattered, effuse, hairy, olivaceous. Mycelium mostly immersed, consisting of branched, septate, smooth, hyaline hyphae. Conidiophores 45–80 lm long ( x = 59 lm, n = 10), 4–5 lm wide ( x = 4.5 lm, n = 10), Distoseptispora palmarum S.N. Zhang, K.D. Hyde & J.K. Liu, sp. nov. Index Fungorum number: IF555431; Facesoffungi number: FoF05090; Fig. 104 Etymology: The epithet reflects the host plant. Holotype: MFLU 18-1588 Colonies on the substratum superficial, effuse, hairy or velvety, black. Mycelium mostly immersed, composed of branched, septate, smooth, pale brown hyphae. Sexual morph Undetermined. Asexual morph Conidiophores 90– 165 lm long ( x = 125 lm, n = 10), 4–7 lm wide ( x = 5.5 lm, n = 10), macronematous, mononematous, septate, unbranched, single or in groups of 2 or 3, erect, Diaporthomycetidae, families incertae sedis Distoseptisporaceae K.D. Hyde & McKenzie Notes: Distoseptisporaceae was established to accommodate a group of sporidesmium-like hyphomycetous taxa, which are phylogenetically distinct from Sporidesmiaceae and characterized by darker conidia with slightly paler rounded apices and indeterminate length, and especially with most species having relatively short conidiophores (Su et al. 2016). Subsequently, a series of new species (Hyde et al. 2016; Yang et al. 2018b; Luo et al. 2018) were introduced to this monotypic family, and Yang et al. (2018b) provided an emendation of Distoseptispora according to much longer, percurrently proliferating conidiophores and euseptate conidia. Two new taxa are introduced to this monotypic family herein, namely Distoseptispora dehongensis and D. palmarum, which were found from freshwater and palm habitats, respectively. 123 146 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 103 Distoseptispora dehongensis (HKAS 101738, holotype). a– c Colonies on substrate. d, e Conidiophores. f, g Conidiophores with conidia. h–m Conidia. n Germinated conidium. o, p Colony on PDA (o from above view, p from below view). Scale bars: n = 20 lm, d– m = 10 lm straight or flexuous, smooth, brown, cylindrical, robust at the base. Conidiogenous cells polyblastic, integrated, determinate, terminal, pale brown to brown, cylindrical. Conidia 35–180 lm long ( x = 76.3 lm, n = 32), 7–11 lm wide ( x = 8.6 lm, n = 32), acrogenous, dry, oblong, obclavate, cylindrical or rostrate, elongated, straight or 147 curved, truncate at the base, rounded at the apex, 7–27distoseptate, smooth, greenish black to brown, paler towards the apex, thick-walled. Culture characteristics: Colonies on PDA, 20–25 mm diameter after 21 days at 25 C, mycelium sparse, dark brown to black, reverse concolourous. Material examined: THAILAND, Trat Province, Koh Chang, on rachis of Cocos nucifera (Arecaceae), 27 April 2017, S.N. Zhang, SNT128 (MFLU 18-0588, holotype), ex-type living culture (MFLUCC 18-1446); ibid. (HKAS 102209, isotype). Fig. 104 Distoseptispora palmarum (MFLU 18-1588, holotype). a Colonies on substrate surface. b, c Conidiophores with conidiogenous cells bearing condia. d Conidiophores. e Conidiogenous cells. f–i Conidia. Scale bars: a = 200 lm, b–d, i = 50 lm, f–h = 20 lm, e = 10 lm 123 148 GenBank numbers: ITS: MK085062, LSU: MK079663, RPB2: MK087670, SSU: MK079661, TEF1-a: MK087660. Notes: Distoseptispora palmarum shares similar conidial features to other species of Distoseptispora, but its conidiophores are longer than most Distoseptispora species (Luo et al. 2018; Tibpromma et al. 2018; Yang et al. 2018b; Phookamsak et al. 2019). Distoseptispora palmarum is similar to D. leonensis (: Ellisembia leonensis) (McKenzie 1995; Shenoy et al. 2006; Su et al. (2016), D. suoluoensis and D. guttulata (Yang et al. 2018b). However, D. palmarum is distinct because of its polyblastic conidiogenous cells. Multi-gene analysis based on LSU, ITS, TEF1-a and RPB2 indicates that D. palmarum and D. leonensis constitute a well-supported independent lineage (Fig. 105). However, D. palmarum differs from D. Fig. 105 Phylogram generated from maximum likelihood analysis based on combined LSU, ITS, TEF1-a and RPB2 sequence data for Distoseptisporaceae species and several related families. Related sequences are taken from Hyde et al. (2016), Su et al. (2016), Xia et al. (2017), Yang et al. (2018b) and Luo et al. (2018). Thirty-four strains are included in the combined analyses which comprise 3612 characters (897 characters for LSU, 694 characters for ITS, 946 characters for TEF1-a, 1075 characters for RPB2) after alignment. Cryphonectria parasitica (CMW 7048) (Cryphonectriaceae, Diaporthales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the 123 Fungal Diversity (2019) 96:1–242 leonensis in the length and number of distosepta in the conidia. Hypocreomycetidae O.E. Erikss. & Winka Notes: Eriksson and Winka (1997) introduced this subclass with four orders. In the classification of Eriksson (2006), the subclass Hypocreomycetidae comprised four orders (i.e. Coronophorales, Halosphaeriales, Hypocreales and Microascales) (Maharachchikumbura et al. 2015, 2016). Most of the taxa of the Hypocreomycetidae have light coloured perithecia, nonamyloid apical rings in the asci when apical rings are present and the absence of true paraphyses (Zhang et al. 2006) Glomerellales Chadef. ex Reblova et al. Notes: We follow the updated classification by Wijayawardene et al. (2018a). Bayesian analysis. The best RaxML tree with a final likelihood value of - 15208.000555 is presented. The matrix had 1181 distinct alignment patterns, with 45.45% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.240242, C = 0.262790, G = 0.284086, T = 0.212881; substitution rates AC = 0.955762, AG = 2.559225, AT = 1.264794, CG = 0.896334, CT = 7.954863, GT = 1.000000; gamma distribution shape parameter a = 0.214652. Bootstrap values for maximum likelihood (ML) equal to or greater than 60 and Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are placed above the branches, respectively. The ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue Fungal Diversity (2019) 96:1–242 Plectosphaerellaceae W. Gams et al. Notes: Several asexual morph genera in this family have verticillate conidiophores such as Acrostalagmus Corda and Verticillium Nees. Stachylidium bicolor, the type of its genus, produces erect, roughened, verticillate conidiophores, often with additional verticillate axes emerging from the main stipe; this results in a more complex conidiophore than in other similar genera (Réblová et al. 2011). Acrostalagmus Corda Notes: Based on the phylogenetic analyses of Réblová et al. (2011), the common tropical hyphomycete described and illustrated by Seifert (1985) as Stilbella annulata was revealed as a member of Plectosphaerellaceae and a sister species to Acrostalagmus luteoalbus, the type of the genus. Both Stilbella annulata and Acrostalagmus luteoalbus 149 produce ameroconidia in bright orange to reddish slimy masses; in both species the reddish pigmentation sometimes also colours the phialides. Acrostalagmus annulatus (Berk. & M.A. Curtis) Seifert, Studies in Mycology 68: 186 (2011) Facesoffungi number: FoF05253; Fig. 106 Saprobic on decaying fruit. Sexual morph Undetermined. Asexual morph Hyphomycetous. forming rounded pale reddish brown slimy heads, oval. Conidiophore axis erect, repeatedly branched, pale reddish brown at the base and hyaline at the apex, full-grown conidiophore stipes 4– 5 lm wide at the base, tapering to 3 lm. Conidiogenous cells phialidic 9–26 9 2–3.5 lm, narrowly flask-shaped in the widest part, arising in whorls of 1–5 at several levels along the main stipe and its branches. Conidia 6–8 9 3– Fig. 106 Acrostalagmus annulatus (MFLU 16-0952, new geographical record). a Host fruit. b, c Conidiomata on host surface. d Conidiophores. e–g Conidiogenus cells and conidiophores with conidia. h, i Conidia. Scale bars: d = 20 lm, e–i = 10 lm 123 150 Fungal Diversity (2019) 96:1–242 4 lm ( x = 7.4 9 3.6 lm, n = 30), hyaline, sub-globose to cylindrical, single-celled. Culture characteristics: Conidia germinated on MEA. On MEA colonies are appressed, circular, flat surface, edge entire, first cream then become dark brown and rise in the centre with mycelium, reverse brown reaching 2 cm in 2 weeks at 18 C. Material examined: THAILAND, Chiang Mai Province, Doi Suthep, 22 December 2015, decaying fruits of unknown species, Subashini C. Jayasiri, C 132 (MFLU 16-0952), living culture (MFLUCC 16-0612). GenBank numbers: ITS: MK607614, LSU: MK607617, RPB2: MK607616, SSU: MK607615. Notes: The new isolate of Acrostalagmus annulatus fits with the type description in having erect, roughened, verticillate conidiophores, phialidic conidiogenous cells and oblong-ellipsoidal conidia accumulate in slime (Seifert 1985; Réblová et al. 2011). Acrostalagmus annulatus (MFLUCC 16-0612) forms a sister clade to A. annulatus (DAOM 212126) and an environment sample with high statistical support (100% ML/1.00 BYPP) in Fig. 107. In here based on morphology and phylogenetic affinity introduced a new record of Acrostalagmus annulatus from Thailand. Fig. 107 Phylogram generated from maximum likelihood analysis based on combined ITS and LSU sequence data of selected taxa of Plectosphaerellaceae. Related sequences were obtained from GenBank. Twenty-two strains are included in the analyses, which comprise 1299 characters including gaps. Tree was rooted to Hypomyces aurantius (GJS74 69) and Trichoderma viride (DAOM JBT1003). Tree topology of the ML analysis is similar to the BI. The best scoring RAxML tree with a final likelihood value of - 4999.737002 is presented. The matrix had 345 distinct alignment patterns, with 41.99% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.222188, C = 0.278776, G = 0.279137, T = 0.219898; substitution rates AC = 0.918516, AG = 1.868080, AT = 2.187888, CG = 0.896478, CT = 5.118188, GT = 1.000000; gamma distribution shape parameter a = 0.637862. Maximum likelihood bootstrap (MLBS) values [ 70% and Bayesian posterior probabilities (BYPP) C 0.95% are given near the nodes respectively. The scale bar indicates 0.04 changes. The ex-type strains are in bold and black. The newly generated sequence is indicated in blue 123 Hypocreales Lindau Notes: The order Hypocreales (Hypocreomycetidae) comprise highly diverse species in the tropics, subtropics and temperature regions (Põldmaa 2011) and compirse the families Bionectriaceae, Clavicipitaceae, Cordycipitaceae, Flammocladiaceae, Hypocreaceae, Nectriaceae, Niessliaceae, Ophiocordycipitaceae, Stachybotriaceae and Tilachlidiaceae with 237 genera (Kirk et al. 2008; Maharachchikumbura et al. 2015, 2016) Fungal Diversity (2019) 96:1–242 Cordycipitaceae Kreisel ex G.H. Sung et al. Notes: The family Cordycipitaceae was first described by Kreisel (1969). However, Cordycipitaceae was validly segregated from Clavicipitaceae by Sung et al. (2007) based on morphological and multi-gene phylogenetic analyses. Most of the species in the family are entomogenous and produce stalked, erect, stromatic ascomata or reduced stipes or subiculate stromata (Kepler et al. 2017). Stroma are fleshy or subiculate and pallid or brightly coloured (Sung et al. 2007; Kepler et al. 2017). Kepler et al. (2017) included 11 genera in this family. Wijayawardene et al. (2017a) included 17 genera in this family. Cordyceps Fr. Notes: Cordyceps was validly published by Link (1833) and formally moved to Cordycipitoideae by Sung et al. (2007) with the type species Cordyceps militaris (L.) Fr. Kepler et al. (2017) synonymised Isaria Pers., Microhilum H.Y. Yip & A.C. Rath., Phytocordyceps C.H. Su & H.H. Wang and Evlachovaea B.A. Borisov & Tarasov under Cordyceps using phylogentically-based results. Cordyceps succavus Y.P. Xiao, T.C. Wen & K.D. Hyde, sp. nov. Index Fungorum number: IF555420; Facesoffungi number: FoF05065; Fig. 108 Etymology: The specific epithet refers to the feature of the stromata (inside hollow after mature). Holotype: MFLU 18-1890 Parasitic in an unknown insect buried in the upper 1 cm of soil, forming yellowish to yellow stromata. Sexual morph: Stromata 4–5 cm long, 3–6 mm wide, yellowish to yellow, fleshly, mostly solitary, stipitate, inside hollow after mature. Stipe 2–3 cm long, 3-5 mm wide, fleshly, yellowish to white, clavate, with stromata on the top. Fertile head 1.5–2 cm long, 4–5 mm wide, fleshly, solitary, cylindrical, yellow to yellowish. Stroma 3–3.4 9 2.4– 2.7 mm ( x = 3.2 9 2.6 lm, n = 30) in vertical section, yellow to yellowish. Perithecia 534–655 9 179–278 lm ( x = 594 9 229 lm, n = 30), half-embedded, ampulliform to ovoid, yellow to yellowish, thick-walled, with ostiole on the top. Peridium 23–36 lm wide ( x = 29 lm, n = 60), two layers, textura porrecta in first layer, textura angularis in secondary layer. Asci 486–600 9 3.6–4.9 lm ( x = 543 9 4.3 lm, n = 60), 8-spored, hyaline, narrow cylindrical, with a thick apex. Apical cap 3.1–4.1 9 3.3– 4.5 lm ( x = 3.6 9 3.9 lm, n = 60) lm diam., hyaline. Ascospores 466–594 9 0.9–1.2 lm ( x = 530 9 1.1 lm, n = 60), filiform, smooth, hyaline, septate, breaking into secondary spores. Secondary spores 2.8–4.9 9 0.9–1.2 lm ( x = 3.8 9 1.1 lm, n = 60), cylindrical, smooth, hyaline, one-celled. Asexual morph: Undetermined. 151 Material examined: THAILAND, Chiang Mai Province, Samoeng, on larvae (Lepidopteran), 10 July 2017, Y.P. Xiao, CM0752 (MFLU 18-1890, holotype); ibid. CM0752b, HKAS 96376 (MFLU 18-1891, isotype). GenBank numbers: ITS: MK086060, LSU: MK086062, RPB1: MK084616, RPB2: MK079353, SSU: MK086058. Notes: According to morphology and phylogenetic analysis (Fig. 110), the new collection is closely related to Cordyceps grylli, which was collected from adults of Gryllidae in China (Teng 1936; Mains 1959), but the phylogenetic analyses supports two different species. Cordyceps succavus is characterized by yellow to yellowish, longer stromata, not becoming grayish when dry, yellow ampulliform to ovoid perithecia, longer asci and shorter ascospores (Table 3). Cordyceps grylli has yellow stromata, becoming grayish when dry, with half-embeded, light brown or chestnut brown, ovoid perithecia, cylindrical asci and filiform ascospores breaking into 1-celled secondary spores (Mains 1959) (Table 3). Therefore, we propose Cordyceps succavus as a new species. Blackwellomyces Spatafora & Luangsa-ard Notes: This genus was introduced by Kepler et al. (2017) with two species with ascospores not disarticulating into secondary spores at maturity and with irregularly spaced septa. Blackwellomyces lateris Y.P. Xiao, T.C. Wen & K.D. Hyde, sp. nov. Index Fungorum number: IF555421; Facesoffungi number: FoF05066; Fig. 109 Etymology: The specific epithet refers to the feature of the stroma (unilateral on the stipe). Holotype: MFLU 18-0663 Parasitic in a larvae (Lepidoptera) buried in the upper 2 cm of soil, dark to dark-brown host, forming yellow to yellowish stromata. Sexual morph Thallus within host, without mycelium cover. Stromata 5–6 cm long, 1.5– 2 mm wide, fleshly, mostly branched, yellow to yellowish, stipitate arising from head of the host. Stipe 3.5–4 cm long, 1.5 mm wide, fleshly, yellow to yellowish, clavate, with stromata on the side face. Fertile head 0.5–2.5 cm long, 1.5–2 mm wide, fleshly, solitary or branched, cylindrical, yellow to yellowish, unilateral on the stipe. Stroma 713– 822 9 939–1106 lm ( x = 767 9 1023 lm, n = 30) in vertical section, yellow to yellowish, unilateral. Perithecia 287–343 9 138–184 lm ( x = 315 9 161 lm, n = 30), embedded, ampulliform to ovoid, yellow to yellowish, thick-walled, with ostiole on the top. Peridium 17–29 lm wide ( x = 23 lm, n = 60), three layers, textura porrecta in first layer, textura angularis in middle layer, textura intricata in third layer. Asci 166–223 9 3.7–5.1 lm ( x = 195 9 4.4 lm, n = 60), 8-spored, hyaline, narrow cylindrical, with a thin apex. Apical cap 2.2–3.0 9 3.3– 123 152 Fungal Diversity (2019) 96:1–242 Fig. 108 Cordyceps succavus (MFLU 18-1890, holotype). a Habitat of Cordyceps succavus. b Stromata without infected insect. c Vertical section of stroma. d Vertical sections showing the semi-immersed perithecia. e Perithecium wall. f, g Asci. h Part of ascospores. i Apical cap of asci. j, k Secondary ascospores. Scale bars: c, d = 500 lm, f, g = 200 lm, h = 50 lm, e = 20 lm, i–k = 5 lm 4.4 lm ( x = 2.2 9 3.3 lm, n = 60), hyaline. Ascospores 160–217 9 1.0–1.5 lm ( x = 188 9 1.2 lm, n = 60) cylindrical, smooth-walled, hyaline, without septa, not breaking into secondary part spores. Asexual morph Undetermined. Material examined: THAILAND, Chiang Mai Province, The Mushroom Research Centre, on larvae (Lepidopteran), 12 June 2017, Y.P. Xiao, MRC170617 (MFLU 18-0663, holotype); ibid. MFLU 18-0664, MRC170617b, (HKAS 96376, isotype). GenBank numbers: ITS: MK086059, LSU: MK086061, RPB1: MK084615, RPB2: MK079354, SSU: MK086057, TEF1-a: MK069471. Notes: According to morphological and phylogenetic analysis (Fig. 110), Blackwellomyces lateris is closely related to B. cardinalis and B. pseudomilitaris, and characterized by yellow to yellowish, longer stromata; yellow to yellowish, longer fertile head with stromata in side face; embedded, ampulliform to ovoid perithecia; shorter asci; aseptate ascospores (Table 4). Molecular data indicate that Blackwellomyces lateris has 16 bp in ITS region, 2 bp in nrSSU, 10 bp in nrLSU, 11 bp in TEF1-a, 12 bp in RPB1, 123 This study 1-celled, 2.8–4.9 9 0.9–1.2 Cylindrical 486–600 9 3.6–4.9 15–20 long, 4–5 wide, fleshly, solitary, cylindrical, yellow to yellowish 20–30 long, 3–5 wide Half-embedded, ampulliform to ovoid, yellow to yellowish 534–655 9 179–278 Filiform, multiseptate, 466–594 9 0.9–1.2 Zang and Kinjo (1998) 1-celled, 4–5 9 1 Cylindric, 300–400 9 4 1–2 thick Ovoid, partly embedded (up to one-half), light brown or chestnut-brown 400–700 9 300–350 Filiform, break into fragments References Ascospores (lm) Asci (lm) Perithecia (lm) Stipe (mm) Yellow when fresh becoming grayish Yellowish to yellow 4–5 cm long, 3–6 mm wide Gryllidae (Orthoptera) Unknown insect C. grylli C. succavus Fertile head (mm) Stromata Host Species Table 3 Synopsis of Cordyceps species discussed in the paper 153 Secondary spores (lm) Fungal Diversity (2019) 96:1–242 20 bp in RPB2 that are different from B. cardinalis (OSC 93610), and has 30 bp in ITS region, 17 bp in nrSSU, 10 bp in nrLSU, 35 bp in RPB1 that are different from B. pseudomilitaris (NBRC 101509). Blackwellomyces lateris has good support in the phylogenetic tree and, therefore, we propose it as a new species. Hypocreales genera incertae sedis Sarocladium W. Gams & D. Hawksw. Notes: Sarocladium was erected by Gams and Hawksworth (1975) and S. oryzae selected as the type species. Sarocladium is classified in Hypocreales, family incertae sedis (Maharachchikumbura et al. 2015, 2016). This genus encompasses 22 species according to Index Fungorum (2019). Sarocladium attenuatum was however, confirmed as synonym of S. oryzae (Bills et al. 2004). Therefore, 20 species are accepted. Sarocladium species resembles Acremonium species, but differ in their elongated, solitary phialides, sparsely or repeatedly branched conidiophores, abundant adelophialides and elongated conidia (Giraldo et al. 2015). Colony and conidial morphology can be the primary characters to identify species. Phylogenetic analysis based on a LSU-SSU and LSU-ITS dataset resolves intraspecific identification (Summerbell et al. 2011; Giraldo et al. 2015; Liu et al. 2017). Species of this genus are rice plant pathogens (Saravanakumar et al. 2009), saprobes, mutualistic endophytes (Yeh and Kirschner 2014) and human opportunistic pathogens (Giraldo et al. 2015). Some endophytes even have a potential as biological control agents against plant pathogens (Kelemu et al. 2001). Sarocladium kiliense (Grütz) Summerb, Stud. Mycol. 68: 158 (2011) : Acremonium kiliense Grütz, Dermatol. Wochenschr. 80:774. 1925. Facesoffungi number: FoF05816; Fig. 111 Holotype: CBS 122.29 Culture characteristics: Colonies attaining a diam of 2.6 cm in 12 days at 22 C, entire margin, circular, dense, moist, with radial wrinkles on the surface, pink from above and reverse. Mycelium strands form at the central of colonies from which the phialides are produced. Phialides 20–46 9 1–2 lm ( x = 34 9 1.7 lm, n = 50) phialidic, slender, tapering towards the apex. Conidia 2–6 9 1–2 lm ( x = 4 9 1.5 lm, n = 100), abundant, cylindrical to elliptical, asepate, hyaline, smooth-walled, straight to curved, with round end, tending to gather in slimy head at the tip of phialide. Material examined: CHINA, Yunnan Province, Kunming Institute of Botany, on scarred leaves of an unidentified plant, 10 September 2017. Rungtiwa Phookamsak, Dry culture (HKAS 101476), living culture (KUMCC 18-0031). 123 154 Fungal Diversity (2019) 96:1–242 Fig. 109 Blackwellomyces lateris (MFLU 18-0663, holotype). a Habitat of Blackwellomyces lateris. b Overview of the host and stromata. c Host. d, e Stromata. f Vertical section of stroma. g Vertical sections showing the immersed perithecia. h Perithecium. i–k Asci. l Apical cap of asci. m, n Ascospores. Scale bars: b– d = 1 cm, e = 2 mm, f, g = 200 lm, i–k = 100 lm, m = 50 lm, h, l, n = 20 lm GeneBank numbers: ITS: MK616348, LSU: MK616474. Notes: Our isolate has a close affinity with Sarocladium kiliense (CBS 122.29) based on the analysis of a combined LSU and ITS sequence data (Fig. 112). Comparison of ITS sequence showed 1/519 bp nucleotides differ from the type strain of S. kiliense (CBS 122.29 = MUCL 9724). Sarocladium kiliense is a combined species which initially proposed by Summerbell et al. (2011) based on the phylogenetic analysis of a combined LSU and SSU sequences. The morphological descriptions of Sarocladium kiliense such as strain MUCL 9724 (Perdomo et al. 2011) and strain IR5 (Ebrahimi and Fotouhifar 2016) were given. Our isolates morphologically resemble type strain MUCL 9724 by having cylindrical, smooth, hyaline, conidia with rounded ends which is in slimy heads and dirty white to pale orange colony, but differ in absent of adelophialides 123 Fungal Diversity (2019) 96:1–242 155 Fig. 110 Phylogram generated from maximum likelihood analysis based on combined ITS, SSU, LSU, TEF1-a, RPB1 and RPB2 sequence data for Ophiocordyceps species and Blackwellomyces. Related sequences are taken from Sung and Spatafora (2004), Sung et al. (2007) and Kepler et al. (2017). Fifity-seven strains are included in the combined genes sequence analyses which comprise total 4558 characters (1002 characters for SSU, 716 characters for LSU, 494 characters for ITS, 920 characters for TEF1-a, 570 characters for RPB1 and 856 characters for RPB2) after alignment. Purpureocillium lilacinum (CBS 284.36) and P. lilacinum (CBS 431.87) are used as outgroup taxa. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best sorting RaxML tree with a final likelihood value of - 37381.826537 is presented. The matrix had 1832 distinct alignment patterns, with 29.69% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.244626, C = 0.262958, G = 0.266932, T = 0.225484; substitution rates AC = 1.386955, AG = 3.678399, AT = 1.009103, CG = 1.080113, CT = 7.048074, GT = 1.000000; gamma distribution shape parameter a = 0.209990. Bootstrap values for maximum likelihood and Bayesian equal to or greater than 75 are placed above and below the branches respectively. The newly generated sequences are indicated in bold and blue and chlamydospores. Our isolates also similar as strain IR5 in colonial characters as well as the shape and size of phialides (tapering toward apex, 19–48 9 1–2 lm) and conidia (cylindrical with round end, 3–7 9 1–2 lm), but differ in absenting of chlamydospores (Ebrahimi and Fotouhifar 2016). Historically, Sarocladium kiliense can be an endophytes (strain CanL-10b, China) on Brassica napus (Zhang et al. 2014b), a pathogen such as strain IR5 (Iran) on Malus domstica leaf (Ebrahimi and Fotouhifar 2016), strain CBS 400.52 (England) on Ficus carica (Lombard et al. 2015), strain CBS 122.29 (Germany) on skin infection of man (Herrera et al. 2013), strain FMR 10426 (USA) on blood of human or animal (Irinyi et al. 2015). Sarocladium kiliense have close association with human infections and plant scab, the detailed morphological description is necessary for rapid diagnosis of agent which causes disease (Perdomo et al. 2011). Herewith we introduce our isolate as a new collection of Sarocladium kiliense. Pleurotheciales Réblová & Seifert Notes: Réblová et al. (2016) introduced the order Pleurotheciales with a single family Pleurotheciaceae Réblová & Seifert based on morphological data and phylogenetic analyses. Pleurotheciaceae Réblová & Seifert Notes: Members of the family Pleurotheciaceae share the following morphological characters: perithecial, astromatic, immersed to superficial ascomata, a hamathecium of paraphyses, unitunicate asci, with a non-amyloid apical ring and hyaline or versicolorous, transversely multi-septate ascospores with polar cells hyaline or brown central 123 This study Aseptate, 160–217 9 1.0–1.5 166–223 9 3.7–5.1 Embedded, ampulliform to ovoid, 287–343 9 138–184 Reddish orange to reddish, Orange 15–30 9 0.9–3 Yellow to yellowish, 50–60 9 1.5–2 Lepidoptera Lepidoptera Lepidoptera B. cardinalis B. pseudomilitaris B. lateris Reddish orange to reddish, cylindrical, elliptical to fusiform, 2–9 9 1–4 Orange to red, cylindrical to enlarging apically 4–50 9 0.5–1.5 Yellow to yellowish, clavate, with stromata on the side face 35–40 9 1.5 Yellow to yellowish, cylindrical, elliptical to fusiform, 5–25 9 1.5–2 HywelJones (1994) Multiseptate, 280–390 9 1 290–410 9 5–6 Superficial with one-quarter to one-third immersed, elongate ellipsoid to elongate ovoid, 290–570 9 120–245 Orange, cylindrical, 2–8 9 1.2–4 Sung and Spatafora (2004) 175–330 9 3–5 Embedded, elliptical to fusiform to obclavate, 230–540 9 110–240 Irregularly multiseptate, 160–320 9 1 Asci (lm) Fertile head (mm) Stromata Host Species Table 4 Synopsis of Blackwellomyces species discussed in this paper Stipe (mm) Perithecia (lm) Reference Fungal Diversity (2019) 96:1–242 Ascospores (lm) 156 123 cells. The asexual morph is characterized by macronematous or semi-macronematous conidiophores, loosely fasciculate or aggregated in indeterminate synnemata, and conidiogenous cells producing conidia holoblastically, conidial secession rhexolytic on short denticles or rachis on sympodially extending polyblastic conidiogenous cells and conidia hyaline, brown or versicolorous, septate or aseptate (Réblová et al. 2016). Phaeoisaria Höhn. Notes: Phaeoisaria was introduced by von Höhnel (1909) for a collection on Gigantochloa sp. (Bambusae) and is typified by P. bambusae Höhn. This dematiaceous hyphomycete genus presently includes 19 species that produce indeterminate synnemata with septate or aseptate ellipsoidal, obovoidal, fusiform-cylindrical or falcate conidia formed on a sympodially extending rachis, occurring on decaying wood, plant debris or soil sediments (Sutton 1973; Deighton 1974; Castañeda et al. 2002; Seifert et al. 2011; Mel’nik 2012; Cheng et al. 2014; Liu et al. 2015; Réblová et al. 2016). Réblová et al. (2012) described the genus Pleurotheciella for two new species, P. rivularia and P. centenaria. Phaeoisaria siamensis Jayasiri & K.D. Hyde, sp. nov. Index fungorum number: IF555710; Facesoffungi number: FoF05324; Fig. 113 Etymology: With reference to country (‘Siam’ earlier name of Thailand) where the specimen was collected. Holotype: MFLU 16-0953 Saprobic on decaying fruits. Sexual morph Undetermined. Asexual morph Conidiomata scattered, indeterminate, erect, rigid, superficial, dark brown composed of compact appressed conidiophores. Conidiophores 330– 380 lm high, 20–25(–30) lm ( x = 360 9 24.5 lm, n = 20) wide, macronematous, in synnematous conidiomata, scattered, synnemata subulate or cylindrical, indeterminate, at the base 13–15 lm beneath the fertile portion with conidiogenous cells, composed of medium to dark brown, smooth, septate parallel hyphae, 2–2.5(–3) lm wide, splaying out at the middle to apex. Conidiogenous cells 8–12 9 2–2.5 lm ( x = 10.5 9 2.3 lm, n = 20), integrated, terminal and intercalary, poly-blastic, sympodial, cylindrical or somewhat attenuated toward the tip, pale brown near base, subhyaline at the apex, denticulate, denticles conspicuously cylindrical, 2–2.5 9 1–1.5 lm. Conidia 5–8 9 3–4 lm ( x = 6.2 9 3.5 lm, n = 30), globose to subglobose, hyaline, aseptate. Culture characteristics: Conidia germinated on MEA. Colonies are appressed, circular, flat surface, edge entire, first cream then become dark brown and rise in the centre with mycelium, reverse brown reaching 10 mm in 2 weeks at 18 C. Material examined: THAILAND, Chiang Mai Province, Doi Suthep, 22 December 2015, decaying fruits of Fungal Diversity (2019) 96:1–242 Fig. 111 Sarocladium kiliense (HKAS 101476, new record). a, b Upper and reverse view of culture on PDA after 17 days. c Mycelium strand form at centra of colony. d Phialides arising from mycelium 157 strand. e, f Phialides bearing conidia. g–j Conidia. Scale bars: d = 100 lm, e, f = 15 lm, g–j = 3 lm 123 158 Fungal Diversity (2019) 96:1–242 Fig. 112 Phylogram generated from maximum likelihood analysis based on combined LSU and ITS sequence data for Sarocladium species. Related sequences were referred to Liu et al. (2017). Twentytwo strains are included in the combined genes sequence analyses which comprise total 1086 characters (517 characters for LSU, 569 characters for ITS) after alignment. Acremonium curvulum (CBS 430.66) (Incertae sedis, Hypocreales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree with clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best sorting RaxML tree with a final likelihood value of - 4666.270461 is presented. The matrix had 300 distinct alignment patterns, with 3.01% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.232286, C = 0.271554, G = 0.293432, T = 0.202727; substitution rates AC = 1.861270, AG = 2.197180, AT = 2.893484, CG = 0.997016, CT = 6.880061, GT = 1.000000; gamma distribution shape parameter a = 0.179594. Maximum parsimony analysis of 791 constant characters and 142 informative characters resulted in two equally most parsimonious tree of 153 steps (CI = 0.614, RI = 0.643, RC = 0.395, HI = 0.386). Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 50 are placed above and below the branches respectively. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.90 are in bold. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Meliaceae sp., Subashini C. Jayasiri, C 128 (MFLU 16-0953, holotype), ex-type living culture (MFLUCC 16-0607). GenBank numbers: ITS: MK607610, LSU: MK607613, RPB2: MK607611, SSU: MK607612. Notes: Phaeoisaria siamensis fits with the generic description of Phaeoisaria in having indeterminate synnemata with aseptate ellipsoidal, obovoidal, fusiformcylindrical or falcate conidia formed on a sympodially extending rachis. Phaeoisaria siamensis (MFLUCC 16-0607) forms a sister clade to P. aquatica (MFLUCC 16-1298) with high statistical support (99% MLBS/1.0 BYPP) in Fig. 114. Phaeoisaria aquatica differs from P. siamensis in having compactly and parallelly adpressed conidiophores, with flared conidiogenous cells in the above half (Luo et al. 2018). A comparison of the ITS and RPB2 nucleotides of these two strains reveals 8 (1.6%) and 13 (1.5%) nucleotide differences which justifies the new species following the guidelines of Jeewon and Hyde (2016). 123 Savoryellales Boonyuen, Suetrong, Sivichai, K.L. Pang & E.B.G. Jones Fungal Diversity (2019) 96:1–242 Notes: The order Savoryellales was introduced by Boonyuen and Pang (2011) to accommodate three sexual genera viz. Ascotaiwania Sivan. & H.S. Chang, Ascothailandia Sri-indr., Boonyuen, Sivichai & E.B.G. Jones (now Canalisporium Nawawi & Kuthub.), Savoryella E.B.G. Jones & R.A. Eaton and an asexual genus Canalisporium based on multi-gene phylogenies. It was placed in subclass Hypocreomycetidae (Sordariomycetes) (Boonyuen and Pang 2011), and subsequently raised to subclass Savoryellomycetidae (Sordariomycetes) by Hongsanan et al. (2017) based on phylogenetic analyses and its stem age (267 MYA) reported in Hyde et al. (2017a). This was also confirmed by Dayarathne et al. (2019a). Savoryellaceae Jaklitsch & Réblová Notes: The family Savoryellaceae has a worldwide distribution in freshwater, marine and brackish water habitats (Ranghoo 1998; Abdel-Wahab and Jones 2000; Cai et al. 2003; Jones et al. 2015). The family was reviewed by Dayarathne et al. (2019a) with morphomolecular and molecular clock analyses. It comprises three genera Ascotaiwania Sivan. & H.S. Chang, Canalisporium Nawawi & Kuthub. and Savoryella E.B.G. Jones & R.A. Eaton (Dayarathne et al. 2019a). Canalisporium Nawawi & Kuthub. Notes: Canalisporium was proposed to accommodate Berkleasmium caribense Hol.-Jech. & Mercado, B. pulchrum Hol.-Jech. & Mercado and C. elegans Nawawi & Kuthub. and typified by C. caribense (Hol.-Jech. & Mercado) Nawawi & Kuthub. Twelve species are accepted in the genus, nine of which have molecular data in GenBank. Canalisporium is characterized by muriform conidia and with sexual morph formerly known in Ascothailandia (Sriindrasutdhi et al. 2010). Canalisporium dehongense W. Dong, H. Zhang & K.D. Hyde, sp. nov. Index Fungorum number: IF555407; Facesoffungi number: FoF05058; Fig. 115 Etymology: Name reflects Dehong, from where the species was isolated. Holotype: MFLU 18-1189 Saprobic on decaying wood submerged in freshwater. Sexual morph Undetermined. Asexual morph Conidiomata sporodochial, scattered, punctiform, pulvinate, granular, black, shiny. Mycelium immersed in natural substrate, consisting of branched, septate, thin-walled, smooth, hyaline to pale brown hyphae. Conidiophores up to 65 lm long, micronematous, mononematous, vesiculate, consisted of 1–9 subglobose, smooth, hyaline cells, unbranched, septate, constricted at the septa. Conidiogenous cells 7–16 9 7–14 lm ( x = 12 9 10 lm, n = 15), holoblastic, monoblastic, terminal, determinate, 159 subglobose, ellipsoidal, sometimes swelling to globose, hyaline, smooth, thin-walled. Conidia 20–30 9 12–19 lm ( x = 25 9 16 lm, n = 40), solitary, acrogenous, ellipsoidal to obovoid, muriform, smooth, brown, 1 straight column of vertical septa and 3–5 rows of transverse septa, slightly constricted at the septa, darkened and thickly banded at the septa, canals in the septa obscured by dark pigmentation; basal cell single, cuneiform, sometimes swollen, pale brown to hyaline. Culture characteristics: On PDA, colony circular, 10 mm in 40 days at 25 C, brown to grey from above, black from below, raised, felty wooly, fairly dense. Material examined: CHINA, Yunnan Province, on submerged wood in a small river, 25 November 2017, G.N. Wang, H158D (MFLU 18-1189, holotype), ex-type living culture, (MFLUCC 18-1396), ibid.; H158D (HKAS 101725, isotype), ex-isotype living culture (KUMCC 18-0075). GenBank numbers: ITS: MK051033, LSU: MK051034, SSU: MK051035. Notes: Canalisporium dehongense is morphologically similar to C. caribense (Hol.-Jech. & Mercado) Nawawi & Kuthub. and C. exiguum Goh & K.D. Hyde in having muriform conidia with one longitudinal septum, two cells at the apex and darkened bands at the septa (Goh et al. 1998b; Sri-indrasutdhi et al. 2010). However, it differs from C. caribense by its smaller conidia (20–30 9 12–19 lm vs. 24–51 9 15–29 lm) and from C. exiguum by having more rows of transverse septa (3–5 vs. 2–3). A comparison of nucleotides of C. dehongense with C. caribense SS03839, C. caribense SS03683 and C. exiguum SS00809 also reveals a difference of 94, 101 and 82 bp in the ITS gene region, respectively. The species clusters as a distinct clade within the genus Canalisporium (Fig. 116) with high statistical support (0.99 BYPP). Sordariomycetidae O.E. Erikss. & Winka Chaetosphaeriales Huhndorf, A.N. Mill. & F.A. Fernández Notes: The order Chaetosphaeriales was introduced by Huhndorf et al. (2004) to accommodate Chaetosphaeriaceae Réblová, M.E. Barr & Samuels. Maharachchikumbura et al. (2015) indicated that Chaetosphaeriaceae and Helminthosphaeriaceae Samuels, Cand. & Magni were sister families within Chaetosphaeriales clade. Subsequently, Konta et al. (2017) introduced another two families, Leptosporellaceae Konta & K.D. Hyde and Linocarpaceae Konta & K.D. Hyde into Chaetosphaeriales based on LSU and ITS sequence data. Besides these four families, Chaetosphaeriales also comprises some genera incertae sedis, e.g. Caudatispora J. Fröhl. & K.D. Hyde, Erythromada Huhndorf, A.N. Mill., F.A. Fernández & 123 160 Fungal Diversity (2019) 96:1–242 Fig. 113 Phaeoisaria siamensis (MFLU 16-0953, holotype). a Host fruit. b, c Conidiomata on host surface. d Conidiophore. e–g Conidiogenous cells arranged on conidiophores. h Conidia. Scale bars: d = 100 lm, e–g = 10 lm Lodge, and Lasiosphaeriella Sivan. (Maharachchikumbura et al. 2015, 2016). Chaetosphaeriaceae Réblová, M.E. Barr & Samuels Notes: Thirty-eight genera are presently accepted in Chaetosphaeriaceae (Maharachchikumbura et al. 2015, 2016; Wijayawardene et al. 2018a; Yang et al. 2018a). Recent studies are those of Liu et al. (2016), Lu et al. (2016), Luo et al. (2016, 2019), Wei et al. (2018) and Yang et al. (2016, 2018). 123 Sporoschisma Berk. & Broome Notes: Most Sporoschisma species have been reported from freshwater habitats (Goh et al. 1997; Ho et al. 2002; Luo et al. 2016, 2019; Yang et al. 2016a). Thirteen species are accepted within this genus (Luo et al. 2016, 2019). Sporoschisma resides in Chaetosphaeriaceae (Chaetosphaeriales) (Maharachchikumbura et al. 2015, 2016). Sporoschisma chiangraiense N.G. Liu & K.D. Hyde, sp. nov. Index Fungorum number: IF555386; Faceoffungi number: FoF04950; Fig. 117 Fungal Diversity (2019) 96:1–242 161 Fig. 114 Simplified phylogram showing the best RAxML maximum likelihood tree obtained from the combined SSU, ITS, LSU and RPB2 matrix of 55 taxa of the orders Pleurotheciales, Savoryellales, Conioscyphales and Fuscosporellales. The matrix comprised 3920 characters including alignment gaps. The tree was rooted to Leotia lubrica (AFTOLID 1) and Microglossum rufum (AFTOL-ID 1292). The best scoring RAxML tree with a final likelihood value of - 32124.131818 is presented. The matrix had 1958 distinct alignment patterns, with 38.37% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.237765, C = 0.258447, G = 0.293741, T = 0.210047; substitution rates AC = 1.443319, AG = 2.967467, AT = 1.388359, CG = 1.234302, CT = 7.043503, GT = 1.000000. MLBS equal to or greater than 70% (first set) and BYPP values equal to or greater than 0.95 (second set) are given at the nodes. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Etymology: The name refers to the place that fungus was collected, Chiang Rai, Thailand. Holotype: MFLU 18-1722 Saprobic on submerged decaying wood. Sexual morph Undetermined. Asexual morph Colonies on natural substrate effuse, black, hairy, with long chains of conidia. Mycelium immersed, composed of hyaline, guttulate, branched, 1.5–3 wide hyphae. Setae 85–150 9 3.5–8 lm, scattered or in groups mixed with conidiophores, capitate, sometimes surrounded by hyaline mucilage at the swollen apex, smooth, median brown, paler towards the apex, straight or flexuous, septate. Conidiophores 100–220 lm long, 5–8 lm wide below venter, 11–13.5 lm wide above, and 13–16 lm wide at venter, macronematous, mononematous, smooth, dark brown to black, paler at the torn apex, straight or slightly flexuous, solitary or in groups of 123 162 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 115 Canalisporium dehongense (MFLU 18-1189, holotype). a, b Colonies on submerged wood. c–i Conidia with conidiogenous cells. j, k Conidia. l Conidiophore. m, n Colony on PDA (m from above view, n from below view). Scale bars: c, d, g, h, l = 20 lm, e, f, i–k = 10 lm 2–3, arising from dark brown to black bulbous base, composed of a cylindrical stipe and a swollen venter with a long cylindrical neck, erect, sometimes proliferating percurrently. Conidiogenous cells monophialidic, percurrent, integrated, brown, lageniform, frayed at the apex. Conidia 15.5–30 9 8.5–12.5 lm ( x = 23.8 9 10.7 lm, n = 30), formed in chains, cylindrical, aseptate, hyaline, verruculose, smooth-walled, with a conspicuous guttule at centre when young, 1-septate, brown, with conspicuous, circular guttules in each cell when mature, conspicuously darkened and slightly constricted at the septa, thickened at both ends (Fig. 117). Culture characteristics: Conidia germinating on water agar media within 48 h. One or two germ tubes produced randomly from conidium wall. Colonies growing on PDA white, irregular, circular, edge entire, mycelium fluffy, dense in the centre, becoming sparse at the edge. Material examined: THAILAND, Chiang Rai Province, Muang District, Ban Nang Lae Nai, on decaying wood submerged in a freshwater stream, 6 March 2018, N.G. Liu, CR066 (MFLU 18-1722, holotype), ex-type living culture (MFLUCC 18-0703). GenBank numbers: ITS: MH883032, LSU: MH883030, SSU: MH883048. Notes: Sporoschisma chiangraiense is phylogenetically related to S. aquaticum and S. palauense (Fig. 118). Comparisons of ITS sequences showed that there are 106 bp differences with gaps of total 558 bp between S. chiangraiense and S. aquaticum, and 58 bp differences with gaps of total 563 bp between S. chiangraiense and S. palauense. Sporoschisma chiangraiense differs from S. aquaticum and S. palauense in having 1-septate conidia when mature, while the latter two are 3-septate, and 1–3septate respectively (Luo et al. 2016; Yang et al. 2016a). Conidia of S. chiangraiense have conspicuous guttules, a feature absent in S. aquaticum. Ellisembia Subram., Proc. Indian Natn Sci. Acad., Part B. Biol. Sci. 58(4): 183 (1992) Notes: Ellisembia, based on E. coronata, was segregated from the widely circumscribed Sporidesmium by Subramanian (1992) and delimited to species with brown, distoseptate conidia and conidiophores with none or percurrent and irregular extension. It comprises saprobic species on wood, bamboo culm or plant remnants in terrestrial and freshwater habitats. Based on the evidence from molecular DNA data, the difference between 163 euseptate vs. distoseptate conidia among Sporidesmiumlike species does not seem significant (Su et al. 2016; Yang et al. 2018). The genus Ellisembia is polyphyletic, it comprises 65 species but only some of them could be positioned in the system based on DNA sequence data. Currently, members of Ellisembia are placed in four clades in Sordariomycetes, i.e. in Chaetosphaeriaceae under the name Ellisembia, Distoseptisporaceae as Distoseptispora, Sporidesmiaceae as Sporidesmium and Xylariales as Ellisembia calyptrata (Réblová and Winka 2001; Shenoy et al. 2006; Su et al. 2016; this study). Until the systematic positions of the type species of Ellisembia and Sporidesmium are revealed, we recommend using the name Ellisembia for species grouped in a strongly supported monophyletic clade in Chaetosphaeriaceae, experimentally linked with sexual morphs having multiseptate, versicolorous ascospores in asci with a non-amyloid apical annulus, persistent paraphyses and immersed ascomata. Ellisembia aurea Réblová & J. Fourn., sp. nov. MycoBank number: MB827909; Facesoffungi number: FoF05783; Figs. 119, 120 Etymology: Aureus (L) golden, referring to the colour of the golden-yellow pigment covering necks and surrounding substrate. Holotype: PRA-14744 Saprobic on decaying wood partly submerged in freshwater. Sexual morph Ascomata immersed with protruding necks or becoming superficial, solitary or in small groups. Venter 600–750 lm diam., 320–400 lm high, subglobose, upright or decumbent. Neck central, 170–200 lm wide, 300–600(–1000) lm long, upright or decumbent, cylindrical, sometimes tapering towards the apex, with a distinct pore at the apex. Neck often basally swollen up to 320–400 lm, conical, surrounded by a dark brown to black stroma formed of densely interwoven, dark brown, septate, thick-walled hyphae 2.5–3.5 lm wide of textura intricata. Stroma covered by erect, dark brown conidiophores. Neck covered irregularly by a layer of golden-yellow granules which disappear with age; granules of the same pigment are also staining the surrounding substrate. Ostiole periphysate. Ascomatal wall leathery, 20–27 lm, two-layered; outer layer consisting of brown, thick-walled, polyhedral cells with opaque walls of textura angularis; towards the interior cells more flattened of textura prismatica. Inner layer consisting of several rows of thin-walled, hyaline, flattened cells. Paraphyses hyaline, branched, anastomosing, septate, 2.5–3.5 lm wide, tapering to ca. 2.0 lm. Asci (165–)170–205 9 23–25.5(–27.5) lm (mean ± SD = 190.4 ± 14.2 9 24.6 ± 1.6 lm), (150–)154–175 lm (mean ± SD = 163.1 ± 8.5 lm) long in the sporiferous part, truncate at the apex, cylindrical to clavate, stipitate; with 8 obliquely uniseriate or biseriate ascospores; apical 123 164 Fungal Diversity (2019) 96:1–242 Fig. 116 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU and SSU sequence data for Canalisporium species and several related genera. Related sequences are taken from Sri-indrasutdhi et al. (2010) and Boonyuen and Pang (2011). Twenty-five strains are included in the combined sequence analyses which comprise total 2757 characters after aligned. Pleurothecium semifecundum (CBS 131271) (Pleurotheciaceae, Pleurotheciales) is used as the outgroup taxon. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best scoring RaxML tree with a final likelihood value of - 11890.841233 is presented. The matrix had 828 distinct alignment patterns, with 10.21% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.225725, C = 0.259821, G = 0.307021, T = 0.207433; substitution rates AC = 1.748411, AG = 3.032505, AT = 1.758208, CG = 1.095962, CT = 6.405200, GT = 1.000000; gamma distribution shape parameter a = 0.244920. Bootstrap values for maximum likelihood (ML) equal to or greater than 70 and Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are placed above the branches, respectively. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue annulus 6.5–7.5(–8) lm wide, 2.0–2.5 lm high, nonamyloid. Outer wall of the ascus fragile, often rupturing basally resulting in a prolonged stipe and apically causing a release of the apical content with the annulus. Ascospores (33–)36–38(– 40) 9 9.5–11.5(– 12) lm (mean ± SD = 37.1 ± 2.0 9 10.6 ± 0.6 lm), ellipsoidal, sometimes inequilateral, 5(–7)-septate, not constricted at the septa, with a large globule in each cell obscuring the septa, versicolorous prior to discharge, middle cells brown to olivaceous brown with hyaline polar cells, smooth-walled, without mucilaginous sheath or appendages. Asexual morph Conidiophores macronematous, mononematous, 92–155 lm long, 6.5–7.5 lm wide, unbranched, upright or slightly bent, seldom with a percurrent extension, septate, dark brown, paler and tapering towards the tip, (3.5– )5.0–5.5 lm wide at the apex, truncate. Conidia holoblastic, secession schizolytic, (65–)70–97(–103) 9 11–16(– 17.5) lm (mean ± SD = 85.0 ± 10.2 9 14.8 ± 2.3 lm), (3.5–)4.5–5.0 lm wide at the base, obclavate to fusiform to lanceolate, truncate at the base, tapering apically with a terminal extension which might be up to 26 lm long, brown to reddish brown, with the darkest basal cell, paler at the apex, 11–13(–15)-distoseptate, smooth-walled. Culture characteristics: Colonies on Modified Leoniańs Agar (MLA) 7–8 mm diam. after 28 d, convex, circular to slightly irregular. Aerial mycelium abundant, lanose, floccose, partially decumbent, colony yellow, surrounded by a yellow-orange pigment diffused into the agar; margin entire to slightly undulate; reverse yellow-orange. Sporulation after 60 d; conidiophores arise at the margin of the colony. Colonies on potato-carrot agar (PCA) 12–13 mm diam. after 28 d, convex, circular. Aerial mycelium abundant around the centre and on the inoculation block, velvety, loose to almost cobwebby towards the margin, colony yellow-orange, surrounded by a yellow zone of submerged growth and deep yellow pigment diffused into the agar; margin filiform; reverse yellow. Sporulation absent. Vegetative mycelium hyaline, some hyphae pale brown, 1.5–2.5 lm wide. Conidiophores 24–40 lm long, 6–10 lm wide in the broadest part, seldom macronematous, mostly semimacronematous reduced to a conidiogenous cell sometimes with 1–2 supporting cells, conidiogenous cell 123 Fungal Diversity (2019) 96:1–242 Fig. 117 Sporoschisma chiangraiense (MFLU 18-1722, holotype). a, b Colonies on natural substrates. c, d Conidia and conidiophores with setae. e Conidiophores and conidia. f Portion of phialide 165 producing conidium. g–i Conidia. j Germinated conidium. Scale bars: a = 200 lm, b = 100 lm, c–e = 50 lm, f–h = 25 lm, i, j = 20 lm 123 166 Fungal Diversity (2019) 96:1–242 Fig. 118 Phylogram generated from maximum likelihood analysis based on combined LSU and ITS sequence data representing Chaetosphaeriaceae. Related sequences are taken from Liu et al. (2016), Lu et al. (2016), Luo et al. (2016) and Yang et al. (2016a). Thirty-five strains are included in the combined analyses which comprise 1635 characters (1028 characters for LSU, 607 characters for ITS) after alignment. Lasiosphaeria ovina (SMH4605) (Lasiosphaeriaceae, Sordariales) is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of - 8025.509056 is presented. The matrix had 607 distinct alignment patterns, with 24.38% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.228162, C = 0.260607, G = 0.308790, T = 0.202441; substitution rates AC = 1.613577, AG = 2.168568, AT = 1.347260, CG = 0.833554, CT = 8.019859, GT = 1.000000; gamma distribution shape parameter a = 0.534284. Maximum parsimony analysis of 1148 constant characters and 117 informative characters resulted in nine equally most parsimonious trees (CI = 0.591, RI = 0.751, RC = 0.443, HI = 0.409). Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 75 are placed above and below the branches respectively. Branches with Bayesian posterior probabilities (BYPP) equal or greater than 0.95 are in bold. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue swelled in the middle, dark brown, with a cup-shaped percurrent extension. Conidia (95–)105–128(– 137) 9 11.5–13.5(–16.5) lm (mean ± SD = 124.0 ± 11.8 9 12.9 ± 1.5 lm), 3.5–4.5(–5) lm wide at the base, 123 Fungal Diversity (2019) 96:1–242 long-fusiform, truncate at the base, tapering apically, brown, with the darkest cell at the base, paler towards the apex, (12–)15–18-distoseptate, smooth-walled, apex sometimes extended into a hyaline filiform appendage 56–85 lm long, 2.0–2.5 lm wide, conidia with an appendage 186–230(–268) lm long (mean ± SD = 221 ± 36.5). Material examined: FRANCE, Ariège, Rimont, La Maille brook, 42.97706, 1.307167, ca. 550 m a.s.l., on partly submerged wood of a decorticated branch of Sambucus nigra, 4 August 2017, J. Fournier, J.F. 17058 (PRA14744, holotype), ex-type living culture (CBS 144403). GenBank numbers: ITS: MH836375, LSU: MH836376. Notes: Conidia that are formed on nature substrate are shorter, usually fusiform to lanceolate, tapering apically with a short terminal extension and with less distosepta compared to conidia formed in MLA culture, which are mostly long-fusiform, wider at the distal end, which sometimes extends into a hyaline filiform appendage. The conidiophores formed in vivo are macronematous and cylindrical compared to usually semimacronematous and swollen conidiophores with a frequent percurrent extension formed in vitro. In the ITS-LSU phylogeny (Fig. 121), E. aurea is positioned in Chaetosphaeriaceae in a strongly supported clade together with Ellisembia brachypus, E. foliiculata and Pyrigemmula aurantiaca. Besides E. foliiculata (Réblová and Winka 2001), E. aurea is another Ellisembia species experimentally linked with a sexual morph with versicolorous ascospores. Ellisembia aurea resembles E. foliiculata in the morphology of immersed ascomata, size of ascospores and morphology of brown, distoseptate conidia but the latter species differs from E. aurea in having flask-shaped ascomata, regularly 7-septate ascospores, shorter and narrower asci and cylindrical to clavate or obclavate, shorter conidia without a rostrum or appendage at the distal end. Hughes and Illman (1974a) reported the characteristic yellowish colour of the substrate surrounding the colonies of E. foliiculata, similar to the golden-yellow pigment released in vitro and found in wood surrounding ascomata of E. aurea. Among Ellisembia species having conidia with a distal rostrate end, E. aurea resembles E. brachypus, which differs in having somewhat shorter conidia (as E. deightonii 50–90 lm fide Ellis 1958, 45–65 lm fide Hughes and Illman 1974b) with fewer, 5–8 distosepta. The conspecificity of E. aurea and E. brachypus (HKUCC 10555, Shenoy et al. 2006) is not supported in our phylogeny. Subramanian (1992) segregated Ellisembia, typified by E. coronata, from the widely circumscribed Sporidesmium to accommodate species with distoseptate conidia and conidiophores with none or percurrent and irregular extension. However, such delimitation based purely on asexual morphological characters is not supported by molecular DNA data (Su et al. 2016; Yang et al. 2018b). Su 167 et al. (2016) resurrected Sporidesmiaceae (Fries 1849) with the single genus Sporidesmium to accommodate taxa with euseptate or distoseptate conidia and accepted Ellisembia as its synonym. However, the type species of neither Sporidesmium nor Ellisembia were included. In the same publication, a new family Distoseptisporaceae and the genus Distoseptispora were introduced (Su et al. 2016), delimiting a group of Sporidesmium-like taxa in having exclusively distoseptate conidia, and two former Ellisembia species were included. However, Yang et al. (2018b) expanded the generic concept of Distoseptispora by including species with euseptate conidia. At present, species formerly classified in Ellisembia are accommodated in four different clades in the Sordariomycetes, i.e. E. adscendens and E. leonensis in Distoseptispora in Distoseptisporaceae, E. bambusicola and E. minigelatinosa in Sporidesmium in Sporidesmiaceae (Shenoy et al. 2006; Su et al. 2016), E. aurea, E. foliiculata and E. brachypus in Chaetosphaeriaceae (Réblová and Winka 2001; Shenoy et al. 2006; this study) and E. calyptrata in Xylariales (Shenoy et al. 2006). Moreover, various sexual morphs have been linked with Ellisembia. Those with hyaline, septate ascospores and superficial or immersed ascomata are currently placed in Distoseptisporaceae, i.e. D. adscendens with Miyoshiella triseptata sexual morph (as Lasiosphaeria, Shoemaker & White 1985), and in Sporidesmiaceae, i.e. S. bambusicola with Miyoshiella fusispora sexual morph (Kawamura 1929) and Sporidesmium thailandense of which only the sexual morph is known (Yang et al. 2018b). The positon of Sporidesmium larvatum with Miyoshiella larvata sexual morph (Reblova 1999) is unknown. Furthermore, Ellisembia sexual morphs with versicolorous ascospores and immersed ascomata were so far exclusively placed in Chaetosphaeriaceae. Despite the proposed synonymy of Ellisembia with Sporidesmium (Su et al. 2016) and also with similar Imicles (Shoemaker and Hambleton 2001; Wu and Zhuang 2005), in the absence of molecular DNA data of their type species, their relationship is unclear and these genera should be retained as separate taxa following Seifert et al. (2011). For the time being, we use the name Ellisembia for taxa attributed to Chaetosphaeriaceae and characterized by pigmented, distoseptate conidia, macronematous conidiophores with or without percurrent extension and multiseptate, versicolorous ascospores in asci with a nonamyloid apical annulus, persistent paraphyses and immersed ascomata. Members of the Ellisembia clade in Chaetosphaeriaceae are congeneric with the morphologically similar Pyrigemmula typified by P. aurantiaca (Magyar et al. 2011). The generic name Pyrigemmula is available and recommended for use to accommodate Ellisembia p. p. and related fungi if its separate position from E. coronata is confirmed in the future. 123 168 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 119 Ellisembia aurea (PRA-14744, holotype). a–c Ascomata. d Vertical section of ascoma (white arrow indicates stroma, black arrow indicates ascomal wall). e Vertical section of the lower part of the neck with visible stroma. f Stroma consisting of interwoven hyphae in detail g Vertical section of the ascomatal wall. h Ascal apex with apical annulus (arrows indicate ruptured outer wall and released ascal content). i, j Asci. k Ascospore. l Paraphyses. m Ascus with immature ascospores. n, o Conidia and conidiophores on nature substrate. Scale bars: a–c = 200 lm, d = 100 lm, e = 50 lm, f = 20 lm, g–o = 10 lm Linocarpaceae Konta & K.D. Hyde Notes: Konta et al. (2017) introduced this family to accommodate Linocarpon Syd. & P. Syd. and Neolinocarpon K.D. Hyde and placed it in Chaetosphaeriales. Senwanna et al. (2018) introduced the novel species, Neolinocarpon phayaoense Senwanna & K.D. Hyde. Neolinocarpon K.D. Hyde Notes: Neolinocarpon was included in Linocarpaceae based on DNA sequence data (Konta et al. 2017) and this was confirmed by Senwanna et al. (2018). Neolinocarpon rachidis Konta & K.D. Hyde Facesoffungi number: FoF05114; Fig. 122 Saprobic on petiole of Cocos nucifera L. Sexual morph Ascomata 320–460 lm high 9 230–290 lm diameter ( x = 415 9 265 lm, n = 10), solitary, deeply immersed, with a central raised, black, globose-subglobose papilla, with a central ostiole. Papilla 157–223 lm high 9 127– 198 lm diameter at the base ( x = 160 9 140 lm, n = 5), black, shiny, with hyaline periphyses. Peridium 14–30 lm wide ( x = 22 lm, n = 10), outer cells merging with the host epidermal cells, composed of dark brown to black cells of textura angularis. Hamathecium comprising numerous, 1–3 lm diameter ( x = 2 lm, n = 20), hyphalike, septate, unbranched, paraphyses, longer than asci. Asci 95–180 9 9–17 lm ( x = 145 9 13 lm, n = 20), 8spored, unitunicate, cylindrical, long pedicellate, with a wedge-shaped, J-, subapical ring. Ascospores 123– 140 9 2–4 lm ( x = 106 9 3 lm, n = 20), parallel in ascus, becoming spiral when mature, filiform, straight or curved, hyaline, aseptate, containing numerous refringent septum-like bands, rounded at the apex, pointed at the base, lacking appendage sheath, smooth-walled. Appressoria 3– 5 9 2–5 lm ( x = 4 9 3.5 lm, n = 20), globose to subglobose, hyaline. Asexual morph Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h and produced appressoria-like structures at each tips of germ tube. Colonies on MEA reaching 2.5–3 cm diameter after two weeks at 25–28 C, white at the edge, brown in the middle with strong radiations outwards. After 3 weeks of incubation, colonies smooth, flat, margin undulate, white to gray in the centre, white at the 169 margin, mycelium becoming red-brown when they produced into media, hyphae septate, branched, smoothwalled. Material examined: THAILAND, Prachaupkhirikhan Province, on dead of petiole, Cocos nucifera (Arecaceae), 30 July 2015, Sirinapa Konta, PJK04i (MFLU 15-2347, isotype), ex-isotype living culture (MFLUCC 15-0814). GenBank numbers: ITS: MK106342, LSU: MK106353, MK106354, SSU: MK106367, MK106368. Notes: Neolinocarpon rachidis was introduced as a new species by Konta et al. (2017) and in this study N. rachidis was found on Cocos nucifera (Arecaceae) in Prachaupkhirikhan Province, Thailand. Appressorium structures were produced by germ tubes indicating that are endophytes. Phylogenetically, N. rachidis strain MFLUCC 15-0814 clustered with N. rachidis MFLUCC 15-0332 with high support (Fig. 123). Therefore, a record of N. rachidis on coconut host is reported. Xylariomycetidae O.E. Erikss & Winka Amphisphaeriales D. Hawksw. & O.E. Erikss. Notes: Eriksson and Hawksworth (1986) introduced the order Amphisphaeriales. However, it was synonymized with Xylariales by Eriksson and Hawksworth (1987). Based on molecular and morphological data, these orders were separated and Amphisphaeriales was resurrected (Senanayake et al. 2015). Currently, this order includes six families, namely: Amphisphaeriaceae, Clypeosphaeriaceae, Bartaliniaceae, Discosiaceae, Phlogicylindriaceae and Pestalotiopsidaceae (Senanayake et al. 2015). Amphisphaeriaceae G. Winter Notes: Winter (1884–1886) introduced the family Amphisphaeriaceae to accommodate the type genus Amphisphaeria. For an account of Amphisphaeriaceae G. Winter see Senanayake et al. (2015). In this manuscript we introduce a new monotypic genus Trochilispora. Trochilispora VP Abreu, AWC Rosado & OL Pereira, gen. nov. Index Fungorum number: IF555484; Facesoffungi number: FoF04859 Etymology: From the family Trochilidae (hummingbirds) and –spora. Type species: Trochilispora schefflerae Abreu VP, Rosado AWC & Pereira OL Asexual morph Associated with leaf spot disease on Schefflera morototoni (Fig. 125a). Conidiomata acervular (Fig. 127c), epiphyllous, scattered and occasionally confluent, subepidermal in origin, erumpent, rounded to oval in outline, unilocular, brown or black, basal stroma thick, of textura angularis, cells thick-walled and almost colourless; lateral walls thick, cells thick-walled, pale brown to brown. Conidiophores cylindrical to subcylindrical, formed in the 123 170 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 120 Ellisembia aurea (CBS 144403, ex-holotype). a Conidium with appendage. b, c Conidia. d, e Conidiophores. f, g Conidiophores with visible percurrent extension. h Conidia formed on aerial mycelium. i, j Colonies on MLA and PCA after 28 d. Scale bars: a–g = 10 lm, h = 100 lm, i, j = 0.5 cm concavity of the conidioma, unbranched, hyaline, smoothwalled. Conidiogenous cells discrete, annellidic with 2 annellations, hyaline, thin- and smooth-walled. Conidia fusiform, straight or slightly curved, 3–4-septate, concolourous, smooth, bearing apical appendage tubular, filiform, single, not centric, unbranched and basal appendage absent; basal cell hyaline to subhyaline, obconic to conic, smooth and thin-walled; 2–3 median cells doliiform, smooth, concolourous, brown, septa darker than the rest of the cell; apical cell hyaline to subhyaline, subconical to hemisphaerical, thin- and smooth-walled. Sexual morph Undetermined. Fig. 121 Phylogram generated from maximum likelihood analysis based on combined ITS and LSU sequence data for Ellisembia aurea and members of Chaetosphaeriaceae. Related sequences are taken from Réblová and Winka (2000, 2001), Réblová (2004), Réblová and Seifert (2008), Schoch et al. (2009), Magyar et al. (2011), Réblová et al. (2011), Yang et al. (2016a) and Vu et al. (2019). Seventeen strains are included in the combined gene sequence analysis, which comprises total 1837 characters including gaps (628 for ITS, 1209 for LSU) and 528 unique characters. Cylindrotrichum clavatum (CBS 125296) and C. oligospermum (CBS 101319) are used as the outgroup taxa. The best RAxML tree with a final likelihood value of 171 Trochilispora schefflerae VP Abreu, AWC Rosado & OL Pereira, sp. nov. Index Fungorum number: IF555485; Facesoffungi number: FoF04860; Fig. 124 Etymology: Name derived from its host genus, Schefflera. Holotype: VIC 44384 Asexual morph Conidiomata acervular, epiphyllous, scattered and occasionally confluent, subepidermal in origin, erumpent, rounded to oval in outline, 49–88 9 79–235 lm diam., unilocular, brown or black, basal stroma thick, of textura angularis, cells thick-walled and almost colourless; lateral walls 3–5 cells thick, cells thick-walled, pale brown to brown. Conidiophores cylindrical to subcylindrical, (8.5–15.5 9 1.5–2 lm), formed in the concavity of the conidioma, unbranched, hyaline, smooth-walled. Conidiogenous cells discrete, annellidic with 2 annellations, (3.5–11.5 9 1.5–3 lm), hyaline, thinand smooth-walled. Conidia fusiform, straight or slightly curved, concolourous, smooth, bearing apical appendage, and basal appendage absent; 3-septate (13–19 9 3.5–5), bearing: [basal cell obconic to conic, hyaline to subhyaline, - 7216.460156 is presented. Estimated base frequencies for the two partitions were as follows: the ITS partition A = 0.214700, C = 0.306167, G = 0.262251, T = 0.216881; substitution rates AC = 4.616604, AG = 5.194154, AT = 5.725980, CG = 3.570615, CT = 13.113689, GT = 1.000000; the LSU partition A = 0.242254, C = 0.240274, G = 0.317255, T = 0.200218; substitution rates AC = 0.504038, AG = 1.058491, AT = 0.427775, CG = 0.207895, CT = 3.855638, GT = 1.000000. Branch support in nodes C 50% maximum likelihood bootstrap support is indicated above branches. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue 123 172 Fungal Diversity (2019) 96:1–242 Fig. 122 Neolinocarpon rachidis (MFLU 15-2347, isotype). a Appearance of ascomata on host substrate. b, c Close up of ascomata. d Yellowish ascospore mass. e Section of ascoma. f Papilla. g Peridium. h Paraphyses. i–k Asci. l J- reaction of apical ring. m– p Ascospores. q Germinated ascospore (appressoria at red arrow). u– y Appressoria. Scale bars: a = 1000 lm, b, d = 500 lm, c = 200 lm, e = 100 lm, f, i–s = 50 lm, g = 20 lm, h = 10 lm, u–y = 5 lm smooth and thin-walled, 2–4 lm long; two median cells doliiform, 8.5–12.5 lm long, smooth, concolourous, brown, septa darker than the rest of the cell (second cell from base brown, 3.5–5.5 lm long; third cell brown, 4.5–7 lm long); apical cell 2–3.5 lm long, hyaline to subhyaline, subconical to hemispherical, thin- and smoothwalled; with 1 tubular apical appendage, arising from the apical crest, not centric, unbranched, filiform, 2–6.5 lm 123 Fungal Diversity (2019) 96:1–242 long; basal appendage absent], or 4-septate (15.5–21 9 4–5), bearing: [basal cell obconic to conic, hyaline to subhyaline, smooth and thin-walled, 2–5 lm long; three median cells doliiform, 10–13 lm long, smooth, concolourous, brown, septa darker than the rest of the cell (second cell from base brown, 4–6 lm long; third cell brown, 2.5–4 lm long; fourth cell brown, 2.5–4.5 lm long); apical cell 2.5–3.5 lm long, hyaline to subhyaline, subconical to hemisphaerical, thin- and smooth-walled; with 1 tubular apical appendage, arising from the apical crest, not centric, unbranched, filiform, 2.5–7.5 lm long; basal appendage absent. Sexual morph Undetermined. Culture characteristics: Colonies cultured on PDA reaching 38 mm diam. after 1 wk at 25 C with a photoperiod of 12 h; regularly margins; with dense aerial mycelium; white; colonies fertile (Fig. 124b). Colonies cultured on MEA reaching 40 mm diam. after 1 wk at 25 C with a photoperiod of 12 h; regularly and submerged margins; with scarce and sebaceous aerial mycelium; pale yellowish; colonies fertile. Material examined: BRAZIL, Minas Gerais, Paraopeba, Floresta Nacional de Paraopeba (FLONA-Paraopeba), on leaves of Schefflera morototoni (Araliaceae), 30 January 2016, V.P. Abreu & O.L. Pereira (VIC 44384, holotype), ex-type living culture (COAD 2371). GenBank numbers: ITS: MH128360, LSU: MH084761, TEF1-a: MH231216, TUB2: MH231215. Notes: Trochilispora is introduced as a new genus based on morphology and phylogenetic support (LSU and ITS sequence data). Based on phylogenetic analyses, Trochilispora schefflerae COAD 2371 grouped in a well-supported clade including Hymenopleella hippophaeicola CBS 140410 (Fig. 125), but different genera can be grouped in the same clade as for example, Morinia and Zetiasplozna; Truncatella and Broomella; among others. Unfortunately, Jaklitsch et al. (2016) did not observed the asexual morph of Hymenopleella hippophaeicola, but the authors cite the Hymenopleella sollmannii species reported by Shoemaker and Müller (1965). The phylogenetic position of the Trochilispora family is still unclear. Trochilispora schefflerae COAD 2371 differs from Hymenopleella sollmannii by having conidia formed in conidiomata acervular with lateral walls 3–5 cells thick of brown hyphae; conidiophores smaller; conidiogenous cells discrete, annellidic with 2 annellations; conidia fusiform, straight or slightly curved, 3–4-septate, with medium brown central cells and hyaline to subhyaline end cells, apical cell with an appendage tubular, filiform, single, not centric, unbranched, not septum and basal cell without appendage basal. Our phylogenetic tree built using LSU and ITS data, and morphological features corroborated that our isolate represents a new genus and a new species belonging to Amphisphaeriaceae (Fig. 125). 173 Sporocadaceae Corda. Notes: Jaklitsch et al. (2016) proposed Sporocadaceae based on morphological observations and phylogenetic analyses of ITS-LSU sequence data. This group of fungi comprises different asexual morph genera which are acervular coelomycetes having hyaline, pale or dark brown, septate conidia. The type genus is Seimatosporium Corda (Jaklitsch et al. 2016). Jaklitsch et al. (2016) accepted Bartalinia, Monochaetia, Neopestalotiopsis, Pseudopestalotiopsis, Pestalotiopsis, Seiridium and many other conidial appendage bearing asexual morph genera (Jaklitsch et al. 2016; Wijayawardene et al. 2017a, 2018a). Wijayawardene et al. (2018a) listed 22 genera in this family. Pestalotiopsis Steyaert Notes: Based on conidial features, Steyaert (1949) divided Pestalotia De Not., into three genera, Pestalotia De Not., Pestalotiopsis Steyaert, and Truncatella Steyaert. Species with 5-celled conidia (4-septate) was grouped within Pestalotiopsis. Based on a multilocus phylogenetic and morphological analysis of pestalotiopsis-like species, Maharachchikumbura et al. (2014) divided the complex into three genera: Pestalotiopsis Steyaert, Neopestalotiopsis Maharachch., K.D. Hyde & Crous and Pseudopestalotiopsis Maharachch., K.D. Hyde & Crous based on morphology of conidial median cells and LSU sequence data. Pestalotiopsis is characterized by median cells concolourous, i.e. three pale-pigmented median cells. Pestalotiopsis is a complex genus and has considerable phenotypic diversity (Maharachchikumbura et al. 2014). Thus, the identification is complemented with DNA sequence data and phylogenetic analysis based on the combination of three gene regions (ITS, TUB2 and TEF1a) (Jeewon et al. 2003; Maharachchikumbura et al. 2011, 2012, 2013; Geng et al. 2013). Pestalotiopsis aggestorum F. Liu & L. Cai, Nature Scientific Reports 7: 870 (2017) Facesoffungi number: FoF05991; Fig. 126 Saprobic on dead leaves of unknown host. Sexual morph Undetermined. Asexual morph Conidiomata acervuli, 130–150 lm diam., pycnidial, superficial to semiimmersed, visible as black spots on the host. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 2–4 9 1–2 lm ( x = 2.4 9 1.3 lm), phialidic, discrete, cylindrical, hyaline. Conidia 17–22 9 7–9 lm ( x = 21.0 9 8.2 lm), fusiform, tapering at both ends, 4septate; basal cell 3–5 lm long ( x = 4.1 lm), conic, hyaline and smooth-walled; three median cells together 11– 15 lm long ( x = 13.7 lm), doliiform, light brown, roughwalled; second cell from base 3–5 lm long ( x = 4.4 lm); third cell 3–5 lm long ( x = 4.6 lm); fourth cell 3–5 lm long ( x = 4.0 lm); apical cell 3.2–4.7 lm long 123 174 Fungal Diversity (2019) 96:1–242 Fig. 123 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU, SSU and TEF1-a sequence data for Neolinocarpon species and several closely related genera in Linocarpaceae and all families in Chaetosphaeriales. Related sequences were referred from Konta et al. (2017) and Senwanna et al. (2018). Twenty-eight strains are included in the combined analyses which comprise 3673 characters (666 characters for ITS, 1005 characters for LSU, 1043 characters for SSU, 946 characters for TEF1-a) after alignment. Sordaria fimicola and Gelasinospora tetrasperma (AFTOL-ID 1287) (Sordariaceae, Sordariales) are used as the outgroup taxa. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis and the Bayesian analysis. The best RaxML tree with a final likelihood value of - 14985.889953 is presented. The matrix had 1082 distinct alignment patterns, with 55.45% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.231817, C = 0.269309, G = 0.304861, T = 0.194013; substitution rates AC = 0.889100, AG = 1.588862, AT = 0.616066, CG = 1.035174, CT = 4.437165, GT = 1.000000; gamma distribution shape parameter a = 0.281924. Maximum parsimony analysis of 3673 constant characters and 649 informative characters resulted in two equally most parsimonious tree of 2562 steps (CI = 0.702, RI = 0.690, RC = 0.484, HI = 0.298). Bootstrap values for maximum likelihood (ML) and maximum parsimony (MP) equal to or greater than 50, Bayesian posterior probabilities (BYPP) equal to or greater than 0.9 are placed above the branches respectively. The ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue ( x = 4.0 lm), conic, hyaline and smooth-walled; 2–3 apical appendages 9–17 lm long ( x = 15 lm), tubular, filiform; basal appendage 5–8 lm long ( x = 6.2 lm), single, central, tubular, filiform. Culture characteristics: Colonies on PDA 35 mm diameter after 7 days at 25 C, circular, raised, dense aerial mycelia with fluffy appearance with circular edge, white; reverse white. Material examined: THAILAND, Chiang Rai Province, Mae Fah Luang District, Doi Mae Salong, on dead leaves of unknown host, 23 September 2016, N. I de Silva, NI104 (MFLU 17-0674), living culture (MFLUCC 17-0869). GenBank numbers: ITS: MK334672. Notes: Our strain clusters with the type of Pestalotiopsis aggestorum in the combined ITS, TUB2 and TEF1-a Pestalotiopsis phylogenetic analyses (Fig. 127). 123 Fungal Diversity (2019) 96:1–242 175 Fig. 124 Trochilispora schefflerae (VIC 44384, holotype). a Leaves of Schefflera morototoni in Floresta Nacional de Paraopeba, state of Minas Gerais, Brazil (the arrows indicate the reproductive structures of the fungus). b Colony on PDA after 1 wk at 25 C with a photoperiod of 12 h in the dark in Petri dishes (90 9 15 mm) (COAD 2371). c Cross section of the conidioma. d–g Conidia. Scale bars: c = 50 lm, d–g = 10 lm aggestorum was isolated from leaves of tea (Camellia sinensis) in Yunnan, China. Our strain has fusiform, 4-sepatate conidia with overlap dimensions. However, our strain has smaller apical appendages (9–17) lm than P. aggestorum (18–28 lm; Liu et al. 2017). The current collection of P. aggestorum was isolated from dead leaves of unknown forest plant in Chiang Mai, as a new collection from Thailand. Xylariales Nannf. Castanediellaceae Hern.-Restr., Guarro & Crous Notes: The phylogenetic position based on analysis of LSU sequence data shows it belongs to Xylariales and is distinct from Beltraniaceae. The typical asexual morph features are: conidiophores macronematous, mononematous or aggregated in sporodochia, branched, brown to pale brown. Conidiogenous cells mono or polyblastic, sympodial, discrete, solitary or in whorls, cylindrical to lageniform, hyaline to subhyaline. Conidia are unicellular or septate transversely, cylindrical, fusiform or lunate, and are 123 176 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 125 Phylogram generated from Bayesian Inference analysis based on combined ITS and LSU sequence data for several closely related genera in Amphisphaeriales. Sequence data of ex-type or exepitype cultures are taken from Senanayake et al. (2015) and the closest hits of GenBank database were included in this study. The combined genes sequence analysis included 40 taxa, which comprise 1405 characters (591 characters for ITS, 814 characters for LSU), and outgroup taxon Immersidiscosia eucalypti (MAFF 242781). Bayesian posterior probability are indicated at the nodes, and values equal or greater than 0.95 are in bold. Isolate numbers are indicated after species names. The ex-type or ex-epitype strains are in bold and black. The newly generated sequence is indicated in bold and blue hyaline. In Castanediellaceae there is single genus Castanediella (Wanasinghe et al. 2018a). Castanediella Hern.-Restr., Crous & M.J. Wingf. Notes: The genus Castanediella was introduced by Hernández-Restrepo et al. (2015), with Castanediella acaciae as type species. Species of Castenediella are differentiated on conidiophores and conidia features. Species belonging to this genus are characterized by branched conidiophores and falcate, cylindrical or fusiform, straight to slightly curved, 0–1-septate conidia. Commonly, species are isolated from wood or leaves and 12 species are currently accepted (Costa et al. 2018; Wanasinghe et al. 2018a). While evaluating the diversity of fungi related to Melipona scutellaris substrates (honey, pollen, surface of nest) living Atlantic Forest, Pernambuco, Brazil, a new species was isolated and it is described based on morphological characteristics and its phylogeny is inferred based on ITS/ LSU sequence data. Castanediella meliponae J.E. Felinto-Santos, R.N. Barbosa & N.T. Oliveira, sp. nov. Index Fungorum number: IF555426; Facesoffungi number: FoF05765; Fig. 128 Etymology: ‘‘meliponae’’ referring to substrate related the Melipona scutellaris. Holotype: URM 91974 Sexual morph Undetermined. Asexual morph Colonies mycelium white, velvety, margin regular, slowly growth, good sporulation, on PDA (potato dextrose agar), MEA (malt extract agar) and OA (Oats Meal Agar) at 25 8C in fourteen days. Mycelium hyaline, septate, smooth, 2–5 lm diam. Conidiophores hyaline, erect, 35–100 9 2–3.5 lm diam., 1-2 septate, branched apex, Conidiogenous cells hyaline, cylindrical, defined shape, smooth 2–3 9 6–30 lm diam. Conidia hyaline, solitary, one celled, falcate, straight to slightly curved with subobtuse apex, smooth, 3–4 9 5–10 lm diam. Culture characteristic: Colinies growing on MEA, surface white, reverse brown with white margins, 29–31 lm 177 diam. On OA surface brown centre and white margin, reverse in brown centre and white margin, 30–35 lm diam. On PDA centre of the dark beige colony and white to light beige margin, reverse with white centre and beige light margin, 20–23 lm diam. Exudates and soluble pigment absent in all media. Cultures incubation at 25 8C in 14 days in incubator type Biological Oxygen Demand. Material examined: BRAZIL, Paulista, Pernambuco state, on surface of nest of Melipona scutellaris, native stingless bees from to Brazilian Atlantic Forest, 29 November 2017, J.E. Felinto-Santos, FIBIO 002-B3 (URM 91974, holotype), ex-type living culture (URM 7933). GenBank numbers: ITS: MH992664, LSU: MH988746. Notes: Castanediella meliponae distinguishes of C. cagnizarii, C. hyalopenicillata and C. eucalypti with respect to conidiophores, conidiogenous cells and conidia morphology. The conidiophores of C. cagnizarii and C. eucalypti are pale brown to brown medium while C. meliponae are predominantly hyaline. Castanediella hyalopenicillata has conidiophores prostrate, while C. meliponae is erect. The conidiogenous cells of C. meliponae distinguish of C. hyalopenicillata, C. cagnizarii, and C. eucalypti in form, absence of denticles in apex, colour and absent of scars. C. hyalopenicillata has conidia fusiform with 0-1 septate. Conidia of C. cagnizarii are cylindrical to filiform with 0-1 septate. Conidia of C. meliponae are falcate and one celled. Furthermore, the conidia of C. meliponae are shorter than C. cagnizarii, C. hyalopenicillata and C. eucalypti. In our phylogenetic analyses (Fig. 129) C. meliponae is posicionated close to Synnemadiella eucalypti, however some morphological diferences were observed between those. Synnemadiella eucalypti produces synnemata and conidia granular, ellipsoid, but inequilateral, base truncate, 0.5 lm diam. These features are not observated in C. meliponae (Fig. 128) and take together with ITS and LSU sequences we consider as new species. Diatrypaceae Nitschke Notes: Taxa in Diatrypaceae can be distinguished by perithecial ascomata, usually embedded in a black stroma, cylindric-clavate to clavate, long pedicellate asci and allantoid ascospores (Glawe and Rogers 1984; Rappaz 1987; Mehrabi et al. 2015; de Almeida et al. 2016). Both coelomycetes (Libertella, Cytosporina) and hyphomycetes (Phaeoisaria) have been reported as asexual morphs of Diatrypaceae (Glawe and Rogers 1984; Wijayawardene et al. 2017a,b; de Almeida et al. 2016; Mehrabi et al. 2016). Currently this family comprises 18 genera (Wijayawardene et al. 2018a; Dayarathne et al. 2019b). Diatrypella (Ces. & De Not.) De Not. Notes: We follow the latest treatment and updated accounts of Diatrypella in Senwanna et al. (2017). 123 178 Fig. 126 Pestalotiopsis aggestorum (MFLU 17-0674, new geographical record). a The specimen. b, c Conidiomata on the host. d–f Section of conidiomata. g Conidiogenous cells. h–j Conidia. Key to the related species of Diatrypella 1. Ascospores 4–5 lm long on average…………… ……………………………………Diatrypella major 1. Ascospores 6–10 lm long on average………………2 2. Entorostroma yellowish or olive-green……………..3 2. Entorostroma white………………………………….4 123 Fungal Diversity (2019) 96:1–242 k Germinating conidia. l, m Colony on PDA (l from above view, m from below view). Scale bars: d, e = 50 lm, f, h = 20 lm, i– k = 10 lm, g = 5 lm 3. Asci larger, 120–150 9 15.5–21.5 lm……………… …………………………………………...D. tectonae 3. Asci smaller, 40 9 8–12 lm………………D. frostii 4. Stromata small, up to 2 mm diam.…………………5 4. Stromata larger than 2 mm………………D. vulgaris 5. 1–4 ascomata per stromata, on twigs of Hevea brasiliensis……………………………..…D. heveae Fungal Diversity (2019) 96:1–242 5. 3–4 ascomata per stromata, on seed pods of Delonix regia……………………………………..D. delonicis Diatrypella delonicis R.H. Perera & K.D. Hyde, sp. nov. Index Fungorum number: IF552536; Facesoffungi number: FOF02658; Fig. 130 Etymology: Named after the host genus. Holotype: MFLU 16-1016 Saprobic on dried seed pods of Delonix regia. Sexual morph Stromata 0.65–1.28 mm wide, solitary or gregarious on host, immersed, with 3–4 ascomata immersed in a single stroma, comprising of outer, dark brown to black, small, tightly packed, thin parenchymatous cell layer, with inner white, loosely packed, parenchymatous cells. Ascomata 243–340 lm high, 217–380 lm diam., ( x = 310 9 331 lm, n = 20), perithecial, immersed in stromatic tissues, aggregated, globose to subglobose, sometimes with flattened base, pale to dark brown, carbonaceous, ostioles not prominent. Peridium 16–24 lm wide ( x = 18 lm, n = 20), 2 layered, inner 3–6 layers of hyaline, flattened, elongate, thin-walled cells of textura angularis, outer 4–7 layers of pale to dark brown, flattened, thick-walled cells of textura angularis. Hamathecium comprising 8.2–13 lm wide ( x = 9.8 lm, n = 20), septate, paraphyses, longer than the asci, tapering towards the apex, wider at the base. Asci 95–100 9 14–20 lm ( x = 100 9 18 lm, n = 25), multi-spored, unitunicate, clavate, with narrow, long, thin-walled pedicel, with swollen upper portion, apex rounded, without an visible apical apparatus. Ascospores 6.2–7.4 9 1.4–1.8 lm ( x = 6.7 9 1.6 lm, n = 20), overlapping, pale brown, allantoid to cylindrical, unicellular, with small, fat globules at the ends, smooth-walled, without sheath or appendages. Asexual morph Undetermined Culture characters: Ascospores germinated on PDA within 18 h and germ tube produced from one or both ends of the ascospore. Reaching 6.5 cm within 14 days on PDA, at 25 C, colonies medium dense, circular, flat, with diffuse margin, white, below similar in color, not zonate, not producing pigmentation on PDA media. Material examined: THAILAND, Chiang Rai Province, Muang District, Mae Fah Luang University, in front of S7 building, on dried seed pods of Delonix regia (Boj. ex Hook.) Raf. (Fabaceae), 10 December 2014, R.H. Perera, RHP 75 (MFLU 16-1016, holotype), ex-type living culture (MFLUCC 15-1014); ibid. RHP 77 (MFLU 16-1032). GenBank numbers: ITS: MH812994, LSU: MH812996, TUB2: MH847790 (MFLUCC 15-1014); ITS: MH812995, LSU: MH812997, TUB2: MH847791 (MFLU 16-1032). Notes: Diatrypella delonicis clusters with D. major, D. frostii and D. vulgaris within Diatrypella sensu stricto (Fig. 131). DNA sequences of D. delonicis and putative strain of D. major (1058) differed in 2 nucleotides in the 179 ITS region. DNA sequences of D. delonicis and putative strain of D. frostii (UFMGCB 1917) differed in 10 nucleotides in the ITS region. Sequence data of the TUB2 region of D. major and D. frostii are not available in the GenBank for the comparison. Diatrypella delonicis and extype strain of D. vulgaris (HVGRF03) showed 1 nucleotide difference in the ITS region, and 14 different nucleotides in the TUB2 region. DNA sequences of D. delonicis and extype strain of D. tectonae (MFLUCC 12-0172a) differed in 7 nucleotide positions of the ITS region. DNA sequence data of D. heveae (ex-type MFLUCC 17-0368) differed in 11 nucleotides in the ITS region and 33 in the TUB2. A key to species is provided above including the phylogenetically closely related species and the taxa were described from Thailand, D. heveae and D. tectonae. Peroneutypa Berl. Notes: Peroneutypa is typfied by P. bellula (Desm.) and currently there are 47 estimated species (Wijayawardene et al. 2017a). The asexual morphs for the genus are unknown. The species within Pernoneutypa are saprobes, pathogens and are widely distributed in terrestrial habitats (Lumbsch and Huhndorf 2010; Maharachchikumbura et al. 2015; Shang et al. 2017). Cultures and sequences for the genus are available, but are lacking for the type. Peroneutypa scoparia Carmarán & A.I. Romero (2006) Facesoffungi number: FoF01998; Fig. 132 Holotype: MFLU 19-0623 Saprobic on decayed wood at a mangrove stand. Sexual morph Ascomata 234–347 9 347–439 lm, solitary to scattered, lacking stromatic tissues, deeply immersed, dark brown to black, globose to subglobose, uni to multi-loculate, ostiolate, papillate. Ostioles central, apex composed of brown outer amorphous layer and inner yellow cells of textura porrecta and ostiolar canal filled with periphyses, 195–530 lm high, 330–720 lm wide. Peridium up to 25– 35 lm wide, thick, 3-layered, incorporating occasionally in the host tissues. Paraphyses 1.5–2 lm wide, arising from the base of perithecia, short, thin-walled cells, guttulate, unbranched with apex blunt. Asci 19.5–23 9 3.2–5 lm ( x = 21 9 4.2 lm, n = 20), 8-spored, unitunicate, clavate, long pedicellate, apically rounded, with J-, subapical ring, initially deliquescing. Ascospores 2.8–4.8 9 1.4–1.9 lm ( x = 4.2 9 1.4 lm, n = 40), overlapping, allantoid, hyaline to light brown, aseptate, not constricted at the septum, smooth-walled with one to few small guttules at the both ends, straight to slightly curved. Asexual morph Undetermined. Culture characteristics: Colonies on PDA, reaching 10 cm diam. after 2 weeks at 20–25 C, medium dense, irregular to regular, slightly raised, surface slightly rough, with uneven edge, undulate margin, cottony, colony from above: pale grey at the margin, greenish-grey at the centre; 123 180 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 127 Phylogram of maximum likelihood analysis based on ITS, TUB2 and TEF1-a sequences for Pestalotiopsis species. Related sequences were obtained from Maharachchikumbura et al. (2012, 2014). Neopestalotiopsis saprophyta (MFLUCC 12-0282) is used as the outgroup taxon. Sixty five strains are included in the combined analyses which comprise 1545 characters (565 characters for ITS, 472 characters for TUB, 508 characters for TEF1-a) after alignment. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the maximum parsimony analysis. The best RaxML tree with a final likelihood value of - 10917.490545 is presented. The matrix had 716 distinct alignment patterns, with 12.01% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.234122, C = 0.292180, G = 0.213547, T = 0.260152; substitution rates AC = 1.238816, AG = 3.435200, AT = 1.254551, CG = 1.048645, CT = 4.518698, GT = 1.000000; gamma distribution shape parameter a = 0.803705. Maximum parsimony analysis of 945 constant characters and 384 informative characters resulted in two equally most parsimonious tree of 1643 steps (CI = 0.525, RI = 0.685, RC = 0.360, HI = 0.475). Bootstrap support values for ML (first set) and MP equal to or greater than 50% are given above the nodes. The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue from below: yellow to brown; mycelium greenish-grey, not produced pigmentation on PDA media. Material examined: THAILAND, Phetchaburi Province, Cha-am District, from a dieback diseased marine wood, 11 January 2018, Vin. Kumar, KRB010 (MFLU 19-0623), living culture (MFLUCC 18-1111). GenBank numbers: ITS: MK603519, TUB2: MK101307. Notes: The species Peroneutypa scoparia (MFLUCC 18-1111) is similar to the type of P. scoparia (MFLUCC 11-0478) in Diatrypaceae, based on molecular data and morphology. It is characterized by immersed ascomata with a cylindrical or subconical, periphysate ostiole, apically rounded, clavate asci and allantoid, aseptate, and hyaline to light brown ascospores. Peroneutypa aeriensis has unique ascomatal characters with multiple, short necks Phylogenetically P. scoparia (MFLUCC 18-1111) shares the same clade with P. scoparia (MFLUCC 11-0478) with a close affinity. Our isolate grouped within the Peroneutypa clade with low statistical support. This species is associated to the dieback symptoms. Here we add a TUB2 sequence with our entry. This is the first Peroneutypa species reported from a mangrove habitat (Jones et al. 2015, 2019). Xylariales Nannf., genera incertae sedis Notes: Xylariales have been revised by several recent studies researchers (Tang et al. 2009; Maharachchikumbura et al. 2015, 2016; Senanayake et al. 2015; Wendt et al. 2018; Daranagama et al. 2018; Voglmayr et al. 2018; Wijayawardene et al. 2018a; Doilom et al. 2018). Maharachchikumbura et al. (2016) accepted 24 genera, 181 while Wijayawardene et al. (2018a) listed 44 genera in Xylariales, genera incertae sedis. Fasciatispora K.D. Hyde The most recent account of this genus are by Daranagama et al. (2018) who we follow here. Fasciatispora arengae Konta & K.D. Hyde Facesoffungi number: FoF05113; Fig. 133 Culture characteristics: Ascospores germinated on MEA within 24 h and germ tube was produced from both ends. Colony on MEA, at first whitish, felty, azonate, with fluffy margins, hyphae, septate, branched, and smooth. Material examined: THAILAND, Phang-nga Province, on dead leaflet of Arenga pinnata (Arecaceae), 5 December 2014, S. Konta, PHR07 g (MFLU 15-0301), ex-type living culture (MFLUCC 15-0326). GenBank numbers: ITS: MK120275, MK120276, MK120277, LSU: MK120300, MK120301, MK120302, SSU: MK120304, MK120305, MK120306. Notes: Doilom et al. (2018) introduced Fasciatispora arengae from southern Thailand based on morphology. Phylogenetic analyses indicated that F. arengae clustered with other species in the ‘Fasciatispora’ clade; F. nypae and F. calami (99 MP/1.00 BYPP, Fig. 134). Accordingly, Hyde (1991), Wijayawardene et al. (2018a) and Doilom et al. (2018) assigned Fasciatispora to Amphisphaeriales genera incertae sedis, which was confirmed by Daranagama et al. (2018). Subsequently, several studies transferred Fasciatispora to Xylariaceae (Alias et al. 1994; Hyde 1995a; Hidayat et al. 2007; Hyde and Wong 1999; Kang et al. 1999; Liu et al. 2015; Hyde et al. 2017a). In this study, phylogenetic analysis of combined four gene regions (TUB2, ITS, LSU, and RPB2) showed that Fasciatispora did not form a clade close to Amphisphaeriaceae as was mentioned in Daranagama et al. (2018), it formed a separate between the Graphostromataceae and ‘Hypoxylon’ clades (Fig. 134). Fasciatispora is, therefore, best accommodated in Xylariales genera incertae sedis. Basidiomycota R.T. Moore Agaricomycetes Doweld Agaricomycetidae Parmasto Agaricales Underw. Cortinariaceae R. Heim ex Pouzar Notes: The limits of the family Cortinariaceae remain unknown. Most species in this family are in the genus Cortinarius, with a cosmopolitan distribution and over 2000 described species. Many genera formerly placed in the Cortinariaceae, (e.g., Phaeocollybia, Hebeloma, Galerina) have been transferred to other families in Agaricales (Matheny et al. 2015). Furthermore, the sequestrate genera, Thaxterogaster, Quadrispora, Protoglossum and Hymenogaster pro parte as well as Cuphocybe, Rapacea 123 182 Fungal Diversity (2019) 96:1–242 Fig. 128 Castanediella meliponae (URM 91974, holotype). a Surface of nest of Melipona scutellaris. b (Left to right) Surface and reverse of colony growth on PDA. c Conidiogenous cell and conidia. d Conidiophore and conidia. e Conidia. Scale bars: c–e = 10 lm and species of Rozites, once thought to be genera within the Cortinariaceae, are currently included in the genus Cortinarius (Peintner et al. 2001, 2002). The basidiocarps range from agaricoid to sequestrate, and many have poorly to well–developed veils. The basidiospores are typically ornamented and cinnamon brown in deposit. Cortinarius (Pers.) Gray Notes: Cortinarius species are important ectomycorrhizal fungi and are associated with different trees and shrubs, belonging to the families Fagaceae, Salicaceae, Caesalpiniaceae, Cistaceae, Dipterocarpaceae, Myrtaceae, Rhamnaceae, Rosaceae and Pinaceae, as well as some herbaceous plants in the Cyperaceae and Polygonaceae (Frøslev et al. 2006). The basidiocarps range from agaricoid to sequestrate, and many have poorly to well-developed veils. The basidiospores are typically ornamented and cinnamon brown in deposit. Hundreds of species are still undescribed and below we present two new species belonging to Cortinarius subgenus Telamonia, sect. Safranopedes (Bidaud, Moënne-Locc. & Reumaux) Liimat., Kytöv. & Niskanen from Europe. Cortinarius minusculus Liimat. & Niskanen, sp. nov. Index Fungorum number: IF555637; Facesoffungi number: FoF05970; Fig. 135a 123 Etymology: Name based on a Latin adjective minusculus, meaning rather small. Holotype: K. Liimatainen & T. Niskanen 12-032 (H) Pileus 0.7–1.5 cm wide, at first conical to somewhat convex, later planoconvex, often with an acute umbo, redbrown, paler at the margin, hygrophanous. Lamellae medium spaced, adnexed to emarginated, brown. Stipe 2.5–5 cm long, 0.15–0.3 cm thick, cylindrical, at first very pale greyish brown, covered with silky, white fibrils, later darker, especially at the lower ‘ of the stipe. Context: in pileus red brown, in stipe yellow brown, dark brown at the base when older. Universal veil white, forming often distinct complete and incomplete on the stipe. Odour in lamellae indistinct. Exsiccatae: pileus brown, stipe greyish brown. Basidiospores (7.5–)8–9(–9.5) 9 4.7–5.3 lm, av. = 8.5 9 4.9 lm, Q = 1.63–1.8, Qav. = 1.74, ellipsoid to obovoid-ellipsoid, moderately verrucose, more strongly at apex, somewhat dextrinoid. Lamellar trama hyphae encrusted. ITS sequence (GenBank MK211177, holotype) distinct from other members of C. sect. Safranopedes. With a sister group relationship to C. scotoides and deviating from it by 10 substitutions and indels. Ecology and distribution: In hemiboreal and temperate mixed forests with Betula, Corylus, Populus and Pyrola rotundifolia on mull soil. Known from northern Europe. Fungal Diversity (2019) 96:1–242 183 Fig. 129 Phylogram generated from maximum likelihood analysis based on combined ITS and LSU sequence data for Castanediella species and several closely related genera. Related sequences are taken from Hernández-Restrepo et al. (2017) and Wanasinghe et al. (2018a). Seventeen strains are included in the combined analyses which comprise 1037 characters (557 characters for ITS, 480 characters for LSU) after alignment. Beltraniella endiandrae (CBS 137976) (Amphisphaeriaceae, Amphisphaeriales), Beltrania pseudorhombica (CBS 138003) (Beltraniaceae, Sordariales) and Beltraniopsis neolitseae (CBS 137974) (Incertae sedis) are used as the outgroup taxa. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood value of - 3681.798352 is presented. The matrix had 245 distinct alignment patterns, with 6.62% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.284691, C = 0.205176, G = 0.254615, T = 0.255518; substitution rates AC = 0.678292, AG = 2.785917, AT = 1.405179, CG = 0.578526, CT = 3.212102, GT = 1.000000; gamma distribution shape parameter a = 3.725904. Bootstrap values for maximum likelihood (ML) equal to or greater than 75 are placed above the branches respectively. The newly generated sequence is indicated in bold and blue Material examined: Finland, Uusimaa, Vantaa, Tammisto, Tammisto’s Nature Reserve Area, herb-rich mixed forest with Quercus, Corylus, Betula, Pinus and Populus, 17 September 2012, K. Liimatainen & T. Niskanen 12-032 (H, holotype; K, isotype). GenBank numbers: ITS: MK211177 (loc. cit. 12-031 (H)), MK211178. Notes: Cortinarius minusculus is a small and slender medium brown Telamonia with distinct white veil, indistinct odour, and ellipsoid to obovoid-ellipsoid spores. It grows on mull soil with several different deciduous trees and Pyrola. The species formed a well-supported monophyletic group within C. sect. Safranopedes in our phylogenetic analysis. Cortinarius subscotoides Niskanen & Liimat. sp. nov. Index Fungorum number: IF555638; Facesoffungi number: FoF05971; Fig. 135b Etymology: The name refers to the affinity to C. scotoides. Holotype: K. Liimatainen & T. Niskanen 12-010 (H) Pileus 2–4 cm wide, at first conical, later planoconvex, with an acute umbo, dark brown, paler at the margin, hygrophanous. Lamellae medium spaced, adnexed to emarginated, at first pale greyish brown, later brown. Stipe 2.5–5 cm long, 0.5–0.9 cm thick, cylindrical, at first silky whitish fibrillose, soon brown, especially at the base of the stipe. Context: in pileus dark brown, in stipe brown, darker brown at the base. Universal veil white, rather thin, forming a complete girdle on the middle part of the stipe. 123 184 Fungal Diversity (2019) 96:1–242 Fig. 130 Diatrypella delonicis (MFLU 16-1016, holotype). a Herbarium material. b Appearance of ascomata on host substrate. c Longitudinal section through stroma with ascomata. d Vertical section through ascomata. e Close up of the peridium. f Paraphyses. g–i Asci. j Ascospores. k Germinating ascospore. l, m Colony on PDA (l from above view, m from below view). Scale bars: b = 1 mm, d = 500 lm, e–i, k = 20 lm, j = 10 lm Odour in lamellae indistinct. Exsiccatae: pileus dark brown, stipe greyish brown. Basidiospores 7–8 9 4.8– 5.3 lm, x = 7.5 9 5.1 lm, Q = 1.4–1.6, x = 1.47, lacrymoid to shortly obovoid-ellipsoid, strongly verrucose, somewhat dextrinoid. Lamellar trama hyphae encrusted. ITS sequence (GenBank MK211175, holotype) distinct from other members of C. sect. Safranopedes. With a close relationship to C. scotoides and deviating from it by five substitutions and indels. Ecology and distribution: In hemiboreal and temperate mixed forests with Corylus on mull soil. Known from northern Europe. Material examined: Finland, Varsinais-Suomi, Lohja, Virkkala, Pähkinäniemi, Corylus forest with some Betula, Populus and Picea on calcareous ground, 16 September 2012, K. Liimatainen & T. Niskanen 12-010 (H, holotype; K, isotype). GenBank numbers: ITS: MK211175 (loc. cit. K. Liimatainen & T. Niskanen 12-015 (H)), MK211176. Notes: Cortinarius subscotoides has a dark brown pileus, short but stout stipe, indistinct smell, and lacrymoid to shortly obovoid-ellipsoid, strongly verrucose spores. It grows on mull, calcareous soil at least with Corylus. In our 123 Fungal Diversity (2019) 96:1–242 phylogenetic analysis the species formed a well-supported clade within C. sect. Safranopedes (Fig. 136). Psathyrellaceae Vilgalys et al. Notes: Psathyrellaceae was proposed by Redhead et al. (2001) based on molecular phylogenetic studies. Initially, the family comprised the genera Psathyrella (Fr.) Quél., Lacrymaria Pat., Coprinellus P. Karst., Coprinopsis P. Karst., and Parasola Redhead et al. The latter three genera were segregated from the polyphyletic genus Coprinus Pers. s. lat. in the same paper. Only a few species (Coprinus comatus (O.F. Müll.) Pers. and related taxa) remained in the Coprinus sensu stricto, and the genus was transferred to Agaricaceae Chevall. Subsequently, several smaller genera were added to Psathyrellaceae. According to Kirk et al. (2008), it comprises 12 genera. Coprinopsis P. Karst. Notes: Members of the genus Coprinopsis live as saprotrophs on various substrata such as wood, soil, dung, herbal remnants or burnt ground. Its basidiomes are agaricoid, brittle, and often short lived and/or deliquescent, with blackish to black spore print (Vesterholt 2012). The genus contains around 200 species globally (Kirk et al. 2008). The phylogenetic tree is shown in Fig. 137. Coprinopsis psammophila Mešić & Tkalčec, sp. nov. Index Fungorum number: IF555781; Facesoffungi number: FoF05972; Fig. 138 Etymology: The species is named after its substrate, sandy soil in a desert shrubland. Holotype: CNF 1/6401 Pileus 10–25 mm broad at maturity, subglobose to ellipsoid at first, later obtusely or truncately conical, often with inflexed margin before expanding, finally subapplanate and radially splitting, partially deliquescent, surface finely and shallowly plicate-sulcate except at the centre, pale grey-brownish to light brown at first, later grey to grey-brown, finally black(ish), entirely covered with a delicate, poorly developed, whitish to cream coloured, silky veil. Lamellae free, rather crowded, up to 2.5 mm broad, soon becoming black, partially deliquescent. Stipe 20–70 9 1.5–3.5 mm, cylindrical, mostly with a broadened (bulbous to subcylindrical) base (up to 6 mm wide), with rather long, completely distinctly narrower or gradually narrowing pseudorhiza, central, widely hollow, whitish, dry, minutely fibrillose-floccose, sometimes striate; basal part always buried in the sand, but often also its lower half, or even most of it. Context very thin. Odour and taste not recorded. Spore print black. Basidiospores [300/3/1] (8.5–)8.8–9.8–10.9(–11.2) 9 5.8–6.4–7.2(–7.5) lm (in KOH 2.5%), averages of different basidiomes 9.5–10 9 6.3–6.6 lm, Q = 1.31–1.52– 1.71, av. Q = 1.47–1.56, ellipsoid or suboblong 185 (subcylindrical), more rarely (sub) ovoid, not flattened, with rounded to slightly acute base and very obtuse apex, smooth, slightly to moderately thick-walled (up to 0.6 lm), with wide (2–3.2 lm in KOH, 1.4–2.5 lm in H2O) and central germ-pore, brown-black in KOH, dark red-brown in H2O, opaque, non-amyloid and non-dextrinoid. Basidia 13–28 9 6–10 lm, clavate, 4-spored, thin-walled, hyaline, surrounded by (3–)4–6(–7) hymenophysalides (pseudoparaphyses). Cheilocystidia not observed. Pleurocystidia conical or fusiform, up to ca. 100 lm long and 18–45 lm wide, thin-walled, hyaline, rather abundant (but scattered) to scarce. Pileipellis a cutis, composed of repent, hyaline, thin-walled, 2–20 lm wide hyphae, with narrowest hyphae on the surface. Veil hyphae 1–9 lm broad, occasionally branched, not or only slightly constricted at the septa, thinwalled to slightly thick-walled (up to 0.5 lm), sparsely diverticulate (excrescences pretty wide with rounded tip), hyaline, mostly smooth, sometimes finely to coarsely encrusted. Clamp connections absent. Habit and habitat: In a scattered group (approximately 30 basidiomes at a radius of ca. 10 m), terrestrial, on a sandy soil among sparse herbaceous plants and low bushes in an arid, lowland, Mediterranean desert shrubland. Distribution: known only from the type locality in Libya, Sirte District. Material examined: LIBYA, Sirte District, by Qasr Abu Hadi town, 13.9 km S-SE from Sirte, 31050 5500 N, 16390 5400 E, 46 m a.s.l., on sandy soil in a desert shrubland, 14 November 2008, leg. M. Čerkez (CNF 1/6401, holotype). GenBank numbers: ITS: MK491274, LSU: MK492278. Notes: Morphologically and ecologically, Coprinopsis psammophila is characterised by occasionally branched and sparsely diverticulate veil hyphae (sect. Alachuani (Singer) D.J. Schafer), mostly ellipsoid or suboblong (subcylindrical), smooth basidiospores with very obtuse apex and wide germ-pore, absence of clamp connections, presence of pseudorhiza, often bulbous stipe base, and terrestrial growth on sandy soil. This set of characters clearly differentiate C. psammophila from all other species in the genus. Among all known species in the section Alachuani, only C. alcobae (A. Ortega) Valade also lives on sandy soil, but has no pseudorhiza, and differs in other morphological characters. Furthermore, only the poorly known Coprinus dryophilus Pat. has a rooting stipe, but lives on wood and also differs in other morphological characters. A megablast search of NCBI’s GenBank nucleotide database, using the ITS sequence from our holotype collection of C. psammophila, showed no closely related species. The closest hit was C. verticillata (SchulzWedd.) Redhead et al. (GenBank Number MH861734, Identities = 653/725(90%), Gaps = 23/725(3%)). In phylogenetic tree of Coprinopsis species based on ITS 123 186 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 131 Phylogram generated from maximum likelihood analysis based on combined ITS and TUB2 sequence data of selected taxa. Related sequences were obtained from GenBank. Ninety-five strains are included in the analyses, which comprise 971 characters including gaps. Single gene analyses were carried out and compared with each species, to compare the topology of the tree and clade stability. Tree was rooted with Kretzschmaria deusta (CBS 826.72) and Xylaria hypoxylon (CBS 122620). Tree topology of the ML analysis was similar to BI. The best scoring RAxML tree with a final likelihood value of - 9028.552721 is presented. The matrix had 404 distinct alignment patterns, with 11.19% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.232808, C = 0.250740, G = 0.235523, T = 0.280930; substitution rates AC = 1.126828, AG = 3.047992, AT = 1.477104, CG = 0.932402, CT = 3.736033, GT = 1.000000; gamma distribution shape parameter a = 0.466736. The maximum parsimonious dataset consisted of constant 316, 546 parsimony-informative and 109 parsimony-uninformative characters. The parsimony analysis of the data matrix resulted in the maximum of two equally most parsimonious trees with a length of 2920 steps (CI = 0.407, RI = 0.751, RC = 0.306, HI = 0.593) in the first tree Maximum likelihood bootstrap (ML) values [ 65%) and Bayesian posterior probabilities (PP) [ 0.90% and are given above the nodes. The scale bar indicates 0.2 changes. The ex-type strains are in bold and black. The newly generated sequences are indicated in bold and blue sequences (Fig. 139) C. psammophila is nested near two other species of sect. Alachuani, C. strossmayeri (Schulzer) Redhead et al. and C. picacea (Bull.) Redhead et al. Coprinopsis villosa L. Nagy, Desjardin, Vágvölgyi & Papp, in Nagy et al., Mycologia 105(1): 120 (2013) Index Fungorum number: IF487717; Facesoffungi number: FoF05973; Fig. 140 Pileus up to 20 mm high and 7 mm broad when still unexpanded, narrowly ellipsoid at first, then narrowly conical (never expanded in our collection), plicate-sulcate except at the centre, white at first, then becoming brownish-grey to grey, covered with medium abundant, rather loosely attached, floccose veil, reddish brown all over or light brown at the wider central zone and whitish around. Lamellae free, crowded, white at first, black at maturity, with whitish edge. Stipe up to 32 9 3 mm, cylindrical or somewhat tapering towards the base, central, widely hollow, dry, tomentulose, white. Odour and taste not observed. Spore print black. Basidiospores [100/2/1] 8.6–10.1–11.4 9 6.2–7– 7.8 lm (in KOH 2.5%), averages of different basidiomes 9.6–10.7 9 6.7–7.3 lm, Q = 1.30–1.45–1.57, av. Q = 1.42–1.47, ellipsoid to ovoid in frontal view, often (sub) amygdaloid in side view, not flattened, with rounded to slightly conical base and rounded apex, smooth, thickwalled (1–1.2 lm), with 1.2–1.8 lm wide, central germpore, medium rusty brown in H2O, medium brown in KOH, semi-transparent, non-amyloid and non-dextrinoid. Basidia 17–33 9 8–11 lm, clavate or spatulate with rounded or subcylindrical upper part, 4-spored, thin-walled, hyaline, 187 surrounded by 3–5(–6) hymenophysalides (pseudoparaphyses). Cheilocystidia up to ca. 40 lm in diameter, mostly subglobose or ellipsoid, thin-walled, hyaline, very delicate and easily collapsing in our material. Pleurocystidia 30–65 9 20–45 lm, ovoid, conical, oblong (subcylindrical) or ellipsoid, thin-walled, hyaline, rather abundant but scattered. Veil on pileus composed of two types of elements: (1) ellipsoid to elongated cells, often inflated and/or constricted at septa, arranged in chains forming tufts, up to 90 lm wide, gradually narrowing towards the pointed tuft tip (to only a few lm), not branched or diverticulate, subhyaline to yellow–brown (with intracellular pigment), thin-walled to moderately thick-walled (up to 1 lm), and (2) narrow, 1–6 lm wide, branched, sparsely diverticulate, hyaline, thin-walled to moderately thick-walled (up to 0.8 lm) hyphae laying on the broader ones. Pileipellis a cutis, composed of repent, ellipsoid to elongated, mostly inflated, hyaline, thin-walled to moderately thick-walled (up to 1 lm), 4–50 lm wide hyphae. Clamp connections present and rather abundant in all tissues. Material examined: CROATIA, Zagreb County, 800 m W-NW from Kostanjevac Podvrški (near Samobor), 45490 5100 N, 15350 1700 E, 250 m a.s.l., 13 October 2007, leg. M. Čerkez (CNF 1/4862, new country record). Habit and habitat: On horse dung (Nagy et al. 2013). Our collection was composed of five immature basidiomes (spores matured after sampling) growing from a heap of horse dung mixed with straw. Distribution: Hitherto known from Croatia, Germany, Hungary, Spain, Sweden and USA (Hawaii) (Nagy et al. 2013; Ruiz & Cerdán 2016). GenBank numbers: ITS: MK491273, LSU: MK492277. Notes: ITS sequence obtained from our material of Coprinopsis villosa is identical to the sequence from the type material (GenBank number HQ847031). C. villosa is characterised by fimicolous habitat (horse dung), brown veil on the pileus (at least near the centre), smooth and medium-sized basidiospores, veil composed of two types of elements (broader, unbranched, non-diverticulate cells in chains together with a narrower, branched and sparsely diverticulate cells), and the presence of clamp connections. According to the dominant type of velar elements, it belongs to the section Lanatuli (Fr.) D.J. Schafer. The only other known Coprinopsis species that has a veil on the pileus composed of sausage like elements together with a branched and diverticulate elements is C. candidolanata (Doveri & Uljé) Keirle et al. However, it has a pure white veil, densely diverticulate narrower veil hyphae, no clamp connections, and it occurs on deer and sheep dung. In the protologue of C. villosa, Nagy et al. (2013) wrote that they did not see clamp connections. However, in Croatian, Spanish (Ruiz & Cerdán 2016) and Hawaiian material (Keirle et al. 2004; mentioned in the protologue under 123 188 Fungal Diversity (2019) 96:1–242 Fig. 132 Peroneutypa scoparia (MFLU 19-0623, new record). a–c Appearance of ascomata on host. d Vertical section through ascoma. e, f Ostiole, section through neck with periphyses. h, i Asci. j–m Ascospores. Scale bars: d = 100 lm, e, f = 50 lm, g–m = 5 lm 123 Fungal Diversity (2019) 96:1–242 ‘‘Other specimens examined’’) clamp connections are present. A length/width ratio (Q) of basidiospores is another difference worth mentioning between the original description and the description of Croatian material. Basidiospores reported in original description are more elongated (Q = 1.47–1.57–1.64) than those measured in Croatian material (Q = 1.30–1.45–1.57). C. villosa is reported here as new to Croatian mycobiota. An overview of coprinoid taxa recorded for Croatia until 2000 is given by Mešić & Tkalčec (2003). Cantharellales Gäum. Notes: The order Cantharellales represents most likely the oldest lineage among homobasidiomycetes and equally the oldest lineage in which the ectomycorrhizal mode of nutrition has been acquired. However, this lineage also harbours few lichen-forming, several lichenocolous as well as many saprotrophic and parasitic species (Hibbett et al. 2014). Also referred to as the cantharelloid clade, this lineage typically contains species producing basidia with a variable number of spores depending on the genus. Two sterigmata are found in many Clavulina and Membranomyces, for example, as well as in some rare tropical African Cantharellus, but a variable number from four to eight sterigmata per basidium in a single species is the situation that predominates in Cantharellus, Craterellus, Botryobasidium and Sistotrema. Botryobasidiaceae Julich. Notes: It was first introduced as subfamily Botryobasidioideae of the Corticiaceae in 1958, and was validly published by Parmasto (1968). Later on, Jülich (1982) raised the status of subfamily to the rank of family, as the Botryobasidiaceae. Corticioid genera Botryobasidium, Botryodontia, Botryohypochnus (considered a synonym of Botryobasidium), Candelabrochaete, Suillosporium, and Waitea, were described based mainly on similarities in the their basidiocarp micromorphology. Botryobasidium has smooth spores and is usually 6–8 sterigmata, while Botryophypochnus produces spiny spores and contains 4-sterigmate basidia. Langer (1994) based on molecular data analysis suggested for synonymization and proved Botryobasidium as paraphyletic. Almost all species produce thin and delicate basidiomata with distinct wide hyphae. Many species are associated with asexual morph genera like Haplotrichum, Acladium, etc. The proposed taxon has somewhat similar in morphological features of Acladium, but molecular data strongly supports to be distinct from all other known genera of Botryobasidiaceae. Therefore, this taxon is being proposed as new genus. 189 Neoacladium P.N. Singh & S.K. Singh, gen. nov. Index Fungorum number: IF556212; Facesoffungi number: FoF5687 Etymology: Name refers to its morphological similarity to the genus Acladium. Type species: Neoacladium indicum P.N. Singh & S.K. Singh Saprobic on dead bark of tree of terrestrial habitats. Asexual morph. Colonies effuse, cottony or velvety, rhizoidal, light yellow,powdery, reverse greyish orange. Mycelium mostly semi-immersed or immersed, single or interwoven, light olivaceous. Stroma none. Chlamydospores produced from lateral hyphae, bulbous and inflated, simple to branched, constricted near septa, pigmented, trident shaped, subhyaline to light olivaceous. Setae and hyphopodia absent. Conidiophores semimacronematous, mononematous, fasciculate, dichotomously loosely interwoven, branched, straight or flexuous, subhyaline to light olivaceous, smooth, septate. Conidiogenous cells integrated to terminal sometimes discrete, determinate, acropleurogenous, polyblastic, inflated or sometimes cylindrical, persistent occasionally sessile, denticulate, denticles numerous. Conidia solitary or catenate, dry, variously shaped, sub-globose- globose, clavateobclavate, lenticular, pyriform to ampulliform, smoothwalled, pigmented, sub-hyaline to light olivaceous. Sexual morph Undetermined. Notes: Phylogenetic analysis of ITS and LSU sequence data indicates that Neoacladium is a distinct genus in Botryobasidiaceae, which forms a clade sister to Botryobasidium. Neocladium is close to Acladium, but differs from all other allied genera in having subhyaline to light olivaceous variously shaped conidia, viz. globose to subglobose, clavate, obclavate, lenticular, ampulliform and pyriform, catenate conidia, dentate and phialidic condiogenous cells and presence of abundant trident like pigmented chlamydospores. Neoacladium indicum P.N. Singh & S.K. Singh, sp. nov. Index Fungorum number: IF556211; Facesoffungi number: FoF5688; Figs. 139, 141 Etymology: Specific epithet ‘‘indicum’’ refers to the country of origin. Holotype: AMH 10054 Saprobic on dead bark of Leucaena leucocephala L. Asexual morph Colonies uniformly spread on dead bark, yellow, velvety. Conidiophores semi-macronematous, mononematous, fasciculate, dichotomously branched, loosely interwoven, straight to flexuous, subhyaline to light olivaceous, smooth-walled, septate, 337–752 9 9–13 lm. 123 190 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 133 Fasciatispora arengae (MFLU 15-0301, new record). a, b Appearance of ascomata on host substrate. c Close up of ascomata. d Section of ascoma. e Peridium. f Paraphyses. g–j Asci. k J ? reaction with Melzer’s reagen in apical ring. l Germinated ascospore. m Colony on culture. n–s Ascospores. Scale bars: a, b = 1000 lm, c = 200 lm, d = 100 lm, g–j = 50 lm, e, f = 20 lm, l, n– s = 10 lm, k = 5 lm Fig. 134 Phylogram generated from maximum likelihood analysis based on combined ITS, LSU, RPB2 and TUB2 sequence data for Fasciatispora species and several closely related genera in Xylariales. Related sequences are taken from Voglmayr et al. (2018), Wendt et al. (2018) and Daranagama et al. (2018). One hundred and fiftyseven strains are included in the combined analyses which comprise 7548 characters (1288 characters for ITS, 2397 characters for LSU, 1246 characters for RPB2, 2617 characters for TUB2) after alignment. Microdochiaceae is used as the outgroup taxon. Single gene analyses are carried out and the topology of each tree had clade stability. Tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood 191 Conidiogenous cells integrated to terminal, sometimes discrete, determinate, acropleurogenous, polyblastic, inflated, sometimes cylindrical, persistent, occasionally sessile, denticulate, denticles numerous, 1–2 celled, 0–1 septate, subhyaline to light olivaceous, 7.48–62.92 lm ( x= 20.52 9 6.32 lm, n = 30). Conidia solitary or produced in short chains, variously shaped, globose to sub-globose, valued of - 156397.840016 is presented. The matrix had 4604 distinct alignment patterns, with 59.36% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.239282, C = 0.272288, G = 0.248516, T = 0.239913; substitution rates AC = 1.231573, AG = 3.991205, AT = 1.258243, CG = 0.936670, CT = 5.480741, GT = 1.000000; gamma distribution shape parameter a = 0.408051. Bootstrap values for maximum likelihood (ML) equal to or greater than 50, Bayesian posterior probabilities (BYPP) equal to or greater than 0.9 are placed above and below the branches respectively. Our strains generated sequences are indicated in blue. Branches with 100% MLBS and 1.00 BYPP are shown as blue nodes 123 192 Fungal Diversity (2019) 96:1–242 Fig. 135 Basidiomata and basidiospores (in MLZ). of a Cortinarius minusculus (K. Liimatainen & T. Niskanen 12-032, H). b Cortinarius subscotoides (K. Liimatainen & T. Niskanen 12-015, H). Photographs Kare Liimatainen. Scale bar: a, b (basidiospores) = 10 lm clavate, obclavate, pyriform, lenticular, sometimes broadly ellipsoidal to ampulliform, pigmented, smooth-walled, base truncate, subhyaline to light olivaceous, 10.5– 25.88 9 9.36–16.5 lm ( x = 15.75 9 13.55 lm, n = 30). Sexual morph Undetermined Culture characteristics: Conidia germinating on malt extract agar (MEA). Colonies light yellow (4A4), reaching 4.2 cm diam. in 10 days at 25 C, rhizoidal with irregular margin, surface powdery, reverse greyish orange (5B5). Hyphae septate, unbranched to branched, pigmented, smooth and thin walled, subhyaline to light olivaceous, 5–5.62 lm wide. Material examined: INDIA, Maharashtra, Pune District, on dead bark of Leucaena leucocephala (Fabaceae), 30 July 2017, P.N. Singh; (AMH 10054, holotype), ex-type living culture (NFCCI 4480). GenBank numbers: ITS: MK391496, LSU: MK391493. Notes: The proposed taxon is compared with the asexual genus Acladium in having subhyaline to light olivaceous ampulliform to pyriform conidia produced from tubular to inflated and conidiogenus cells which are persistent, and occasionally sessile. Chlamydospores in the proposed taxon produced laterally and form trident like structure, which are pigmented. Most of these features are absent in Fig. 136 Phylogram resulting from the RaXML (Stamatakis 2014) analysis of ITS regions. Bootstrap values greater than 50% are indicated above branches. The names in blue represent the new species of Cortinarius and the specimens in boldface the type specimens of the species. The tree is rooted with C. section Bovini. The ex-type strains are in bold. The newly generated sequences are indicated in blue 123 Fungal Diversity (2019) 96:1–242 Fig. 137 Maximum likelihood phylogenetic tree of Coprinopsis species based on an ITS1–5.8S–ITS2 DNA sequence alignment and calculated with PhyML software (Guindon et al. 2010). Related sequences are taken from Nagy et al. (2011, 2012, 2013), Raut et al. (2011), Osmundson et al. (2013), Raut et al. (2015), Tibpromma et al. (2017), Phookamsak et al. (2019) and Vu et al. (2019). Forty-three strains are included in the gene sequence analyses with total of 737 characters after alignment. The tree is rooted with Coprinellus xanthothrix (SZMC-NL-3417) and C. domesticus (SZMC-NL-1292). The best PhyML tree with a final likelihood value of -3591.26595 is 193 presented. Estimated base frequencies were as follows: A = 0.23036, C = 0. 0.22632, G = 0.21420, T = 0.32912; substitution rates AC = 1.48357, AG = 3.68147, AT = 1.56733, CG = 0.64802, CT = 4.50774, GT = 1.00000; gamma distribution shape parameter a = 0.570; proportion of invariant is 0.348. The ex-type strains are indicated in bold. The newly generated sequences are indicated in blue. Maximum likelihood bootstrap values greater than 70% are indicated at the nodes. The bar indicates the number of nucleotide substitutions per site 123 194 Fungal Diversity (2019) 96:1–242 Fig. 138 Coprinopsis psammophila (CNF 1/6401, holotype). a, b Basidiomes. c–i Basidiospores. j–l Pleurocystidia. m Veil on the pileus (phase contrast). Scale bars: a = 10 mm, b = 5 mm, c–i = 3 lm, j–m = 10 lm 123 Fungal Diversity (2019) 96:1–242 195 Fig. 139 Neoacladium indicum (AMH 10054, holotype). a Yellow velvety fungal colonies spreaded on outer bark surface. b Stereoscopic surface view of fungal colonies from substratum. c Colony morphology (above view). d Dichotomously branched conidiophores. e Enlarged microscopic view of dichotomously branched conidiophores with intercalary, lateral and terminal dentate conidiogenous cells and conidia. f A detached branch of conidiophore bearing conidiogenous cells and conidia. g Catenate conidia. h–i Trident shaped and inflated chlamydospores. Scale bars: d–i = 20 lm Acladium. However, the conidial shape, length–width are very much variable in N. indicum, which separate it from other reported taxa in this group. Proposed taxon is also distinct from other genera in this group, like Burgoa, Minimedusa, Ceratobasidium, (Rogers 1935; Weresub and LeClair 1971) in having non-bulbuliferous vegetative parenchymatous propagules. During this study the sexual stage Botryobasidium of Acladium was not encountered. The proposed N. indicum (AMH 10054 holotype), is different from other Botryobasidium species based on the 123 196 Fungal Diversity (2019) 96:1–242 Fig. 140 Coprinopsis villosa (CNF 1/4862, new geographical record). a Basidiomes. b Basidiospores. c Veil on the pileus. d Veil on the pileus (phase contrast). e Cheilocystidia (phase contrast). f–h Pleurocystidia. Scale bars: a = 3 mm, c = 100 lm, e–h = 20 lm, b, d = 10 lm sequence analyses. On megablast analysis, ITS sequence of N. indicum showed 80% (524/651) identity and 53 gaps (8%) with B. subcoronatum, 84% (343/410) identity and 27 gaps (6%) with B. simile (GEL2348), and 81% (521/645) identity 54 gaps (8%) with B. intertextum UC2022959. The phylogenetic analysis clearly establishes Neoacladium as a 123 novel genus, and N. indicum as the type species with strong bootstrap values (Figs. 142, 143). Hydnaceae Chevall. = Clavulinaceae Donk, Beih. Nova Hedwigia 1(4): 407 (1970) Fungal Diversity (2019) 96:1–242 = Cantharellaceae J. Schröt., in Cohn, Krypt.-Fl. Schlesien (Breslau) 3.1 (25–32): 413 (1888) [1889] = Sistotremataceae Jülich, Bibl. Mycol. 85: 390 (1982) = Pterygellaceae Jülich, Bibl. Mycol. 85: 386 (1982), = Heteroacanthellaceae P. Roberts, Mycologist 12(4): 147 (1998) = Repetobasidiaceae Jülich, Bibl. Mycol. 85: 388 (1982) Notes: Index Fungorum (http://www.indexfungorum. org) and MycoBank (www.mycobank.org) are still considering Hydnaceae in a different, much narrower sense, whereas recent phylogenetic studies (Moncalvo et al. 2006; Hibbett et al. 2014; Lawrey et al. 2016) have demonstrated that Hydnum L., the type genus of Hydnaceae, belongs in the core lineage of the cantharelloid clade. Hence, the older family name ‘‘Hydnaceae’’ has nomenclatural priority over other family names that could be considered to name the family that groups members of this core clade. The morphological character that seems to support this lineage is the stichic nature of basidia (Pine et al. 1999), although this is only based on observations for very few species. The family is principally comprised of edible species (all Craterellus, Cantharellus and Hydnum). Type genus: Hydnum L., Species Plantarum: 1178 (1753) Cantharellus Adans. Notes: Cantharellus is an ectomycorrhizal genus of edible mushrooms and is reputed for its high culinary and commercial value. The genus has a worldwide distribution and is one of the prominent fungal genera in Africa (De Kesel et al. 2016) and Madagascar (Buyck et al. 2015), with several species being sold along roads and in local markets in large quantities. The genus is infamous for the difficulty of species delimitation and identification (Olariaga et al. 2016). The recent use of DNA sequences for the delimitation of species has more than doubled the number of worldwide accepted species over the past ten years, with numerous species continuing to be described from the African continent (e.g. Buyck et al. 2017; Das et al. 2018a). Sequencing of recent collections from Africa and Madagascar has also allowed for the epitypification of many of the earlier species (e.g. Buyck and Hofstetter 2018; Buyck et al. 2018). Cantharellus goossensiae (Beeli) Heinem., Bull. Jard. Bot. Etat Brux. 28(4): 406 (1958). MycoBank number: MBT385407; Index Fungorum number: IF556561; Figs. 144, 145 Basionym: Hygrophorus goossensiae Beeli, Bull. Société Royale Botanique Belge 61: 99 (1928). Iconography: HEINEMANN (1958, Fig. 42; 1959, pl. XXVII, Fig. 3). Original diagnosis: « Pileo carnoso crasso, convexe, glabro levi, atro-violacea, 5–6 cm lato; stipite crasso solida dein cavo, sulcato, ochraceo, 3–4 9 2 cm; lamellis crassis decurrentibus venoso connexis, ochraceis; sporis 197 ellipsoideis, hyalinis, levis, 8–9 9 4–5 lm; carne ochraceopallida contactu aeris rubescente. » Original description (freely translated from French)‘‘Pileus thick and fleshy, unevenly convex, smooth, glabrous, dark violaceous, 5–6 cm diam.; stipe thick, solid then hollowing, ochraceous, 3–4 9 2 cm; gills thick, decurrent, interveined, yellowish ochre; spores ellipsoid, hyaline, smooth, 8–9 9 4–5 lm; flesh becoming pinkish when cut or exposed.’’ Epitype description, Figs. 144, 145 Basidiomata fleshy and firm, dispersed or solitary. Pileus 82 mm diam., with strongly inrolled margin, at surface dull, tomentose to fibrillose, yellowish brown with olive tinges, slightly paler in the centre (4–5DE series), toward the margin rather dull yellowish to grayish brown, musterd brown (5E6–4, 5F4–5). Hymenophore composed of slightly higher radial gill folds compared to the strongly developed interstitial venation, nearly off-white to pale cream when young, then developing ochraceous tints. Stipe thick and massive, 30 9 11-26 mm, surface splitting in horizontally arranged patches or rims in the upper portion, smooth elsewhere, slightly paler than pileus and developing dark grayish brown to orange brown tinges, not hollowing. Context off-white to pale cream, moderately yellowing, finally turning ferruginous when cut, 6-7 mm thick above the upper stipe-hymenophore transition. Smell faint, but typical and fruity. Taste mild. Spore print insufficient but very pale. Spores ellipsoid to narrowly ellipsoid, (5.8)6.2–6.57– 7.0(7.3) 9 (3.8)4.0–4.38–4.7(5.0) lm, Q = (1.3)1.4–1.51– 1.6(1.7), smooth, hyaline. Basidia narrowly clavulate, measuring mostly 40–55(63) 9 7–8(9) lm, predominantly 4–5spored. Subhymenium filamentous. Cystidia not observed. Pileipellis composed of long, multi-celled hyphal extremities, aggregating together in trichoids or tufts, composed of subcylindrical, thin-walled cells; the terminal cell variable in length, mostly 30–70 lm long, ca 10–20 lm wide and most often wider than the subterminal cells, often slightly constricted in the middle portion, sometimes even clavate, broadly obtuse at the tip, rarely subcylindrical or somewhat narrowing upward. Clamp connections present everywhere. Epitype: CENTRAL AFRICAN REPUBLIC. DzangaSangha Forest Reserve, near Bayanga, close to Bai-Hakou base camp, N2.859934 – E16.467492, under monospecific Gilbertiodendron dewevrei forest, on bare sandy soil, 19 May 2016, 1648/Buyck 16.064 (PC 0125008), epitype is designated here !; Democratic Republic of the Congo. Central African forest district, near Selenge-Lukolela, isolated on the soil of the Gilbertiodendron dewevrei forest, August 1925, M. Goossens-Fontana 372 (BR), holotype. Additional material examined and sequenced: CENTRAL AFRICAN REPUBLIC. Dzanga-Sangha Forest Reserve, near Bayanga, close to Bai-Hakou base camp, N2.859934 – E16.467492, under monospecific 123 198 Fungal Diversity (2019) 96:1–242 Fig. 141 Neoacladium indicum (AMH 10054, holotype). a Branched chlamydospores. b Hyphae and different types of conidia. c Numerous globose to sub-globose conidia. d, e Catenate conidia. f Lenticular conidium. g Pyriform conidium. h, i Clavate conidia. j, k Obclavate conidia with protuberant hila. Scale bars: a–k = 20 lm Gilbertiodendron dewevrei forest, on bare sandy soil, 16 May 2016, 1644/, Buyck 16.032 (PC 0124999); id., 1647/Buyck 16.063 (PC 0125002), 18 May 2016, 1655/Buyck 16.080 (PC 0125003); id., 1660/Buyck 16.087 (PC0125001); id., 1662/Buyck 16.091 (PC0125004); id., 19 May 2016, 1664/Buyck 16.096 (PC0125000); id., 1665/Buyck 16.097 (PC0125007); id., 26 May 2016, 1679/Buyck 16.127 (PC0125005); id., 24 May 2016, Buyck 16.114 (PC0125006). Notes: This species, which was first described as a Hygrophorus (Beeli 1928), seems locally quite common in the Gilbertiodendron dewevrei rain forest. The original description was based on a watercolor sketching a single, very juvenile fruiting body. The latter, therefore, is hardly representative for what this species might look like when fully mature, particularly when considering the impressive variability of the general field habit repeatedly reported already for other chanterelles (Buyck et al. 2016b, 2016c; Das et al. 2018b; Olariaga et al. 2015). Moreover, the original description was very succinct, both for macro- and microscopic features, and sketches a species characterized by a convex, irregularly lobed and dark violet young cap sitting on top of an ochraceous yellow, short and fleshy stipe and having a similarly colored, strongly veined-anastomosing hymenophore. Yet, the feature that has most frequently been emphasized as characteristic for this species (Eyssartier 2001; 123 Fungal Diversity (2019) 96:1–242 199 Fig. 142 Molecular phylogenetic analysis of ITS gene region of Neoacladium indicum (AMH 10054, holotype) by Maximum likelihood method. Sixteen strains are included in the sequence analyses which comprise 668 positions after alignment. Auricularia auricularjudae CBS 7 (JF440710) (Auriculariaceae, Auriculariales) is used as the outgroup taxon. Bootstrap values for maximum likelihood (ML) greater than 50 are placed along the branches, respectively. The evolutionary history was inferred by using the Maximum likelihood method based on the Jukes-Cantor model (Jukes and Cantor 1969). The tree with the highest log likelihood (- 4900.5990) is shown. Evolutionary analyses were conducted in MEGA7 (Kumar et al. 2016). The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue Buyck et al. 2012; De Kesel et al. 2016) is the mention - in the original description - of the unusual pinkish reddening of its context when exposed, as opposed to the yellowing context of most other Central African chanterelles with similar colors. As already previously discussed (Buyck et al. 2012), we think that this pinkish red context is probably a secondary and perhaps unimportant or even illinterpreted character. Very similar examples were recently discussed for two North American species (Buyck et al. 2016b), including C. septentrionalis A.H.Sm., for which the type description equally mentions a characteristic pinkish lilac tinge in the context. Fortunately, however, the microscopic features of C. goossensiae allow to distinguish it rather easily from other tropical African chanterelles as it offers the unique combination of comparatively small, ellipsoid spores and a pileipellis composed of thin-walled, clamped hyphal extremities terminating frequently in rather short but quite wide (11–13 lm diam.), obtusely rounded terminal cells (Heinemann 1958; Buyck et al. 2012) as here indeed illustrated again for the epitype (Fig. 147c). The spores of the holotype are clearly smaller than mentioned in the original description, viz. 5.5–6.72–8 9 3–3.97–5 lm, Q = 1.37–1.71– 2.0 (see Eyssartier (2001)). The small spore size is here confirmed by our collections and is a useful feature to distinguish this species from C. cyanoxanthus (see Buyck et al. (2012)). During a recent field trip in the Central African rain forest, not so far from where this species was originally described, Buyck collected many specimens sharing identical microscopic features with the holotype of C. goossensiae. One of these included a juvenile specimen that corresponded exactly to the published watercolor of the holotype (see Fig. 145a). As a result, we have no doubt about the correctness of our identification as C. goossensiae and we can finally epitypify this enigmatic species and assess its phylogenetic affinities and morphological variability using molecular sequencing. The epitype description is based on the epitype only, but we have obtained (near-) identical sequences from all of our collections, (seven sequences have been used to represent the species in our phylogenetic analysis, Fig. 146). These sequences confirm the impressive field variability of this species, both in size and general color. Variations in size concern not only the stipe, which can exceptionally be more slender and up to 6 cm long, but also the pileus which 123 200 Fungal Diversity (2019) 96:1–242 Fig. 143 Molecular Phylogenetic analysis of LSU of Neoacladium indicum (AMH 10054, holotype) by Maximum likelihood method. Twenty seven strains are included in the sequence analyses which comprise 2176 positions after alignment. Auricularia auricular-judae CBS 7 (JF440710) (Auriculariaceae, Auriculariales) is used as the outgroup taxon. Bootstrap values for maximum likelihood (ML) greater than 50 are placed along the branches respectively. The evolutionary history was inferred by using the Maximum likelihood method based on the Kimura 2-parameter model (Kimura 1980). The tree with the highest log likelihood (- 3161.4740) is shown. Evolutionary analyses were conducted in MEGA7 (Kumar et al. 2016). The ex-type strains are in bold and black. The newly generated sequence is indicated in bold and blue reached 15 cm diam. in one collection). Moreover, C. goossensiae is another example of a species that changes drastically in general appearance between very young and more adult stages, exactly as for the closely related and recently neotypified African C. miniatescens (Buyck et al. 2016a). Very young stages of the latter species are entirely velutinous and dark wine red, while the adult stages develop into nearly smooth, bright yellow basidiomata (for illustrations, see Buyck et al. (2016a)). This is exactly what happens in C. goossensiae where very young stages are dark violet and tomentose, but then tend to become smooth and much paler with age, usually turning brownish to yellow or even pale cream in one collection. The explanation of this metamorphosis lies in both cases with the dense fibrillose to tomentose layer that entirely covers the young cap of both species. This surface layer is of a very different color compared to the color of the underlying cap surface which will be gradually exposed as it expands. Secondly, depending on the intensity of the yellowing of the context (in which we could not detect any pinkish tinges), the overall color of the mature pileus varies greatly, from pale cream over yellow to reddish brown. For somebody acquainted with European chanterelles, some specimens of the African C. goossensiae are reminiscent of C. amethysteus in the field because of a similar, robust stature and identical, violet-purplish tinges on the pileus surface (see Olariaga et al. (2016)). Notwithstanding the fact that both species share the omnipresence of clamp connections in their tissues, they belong in different subgenera (Fig. 146). Indeed, C. goossensiae appears to belong in the exclusively African subg. Pseudocantharellus. The species is strongly reminiscent of C. miniatescens, 123 Fungal Diversity (2019) 96:1–242 201 Fig. 144 Cantharellus goossensiae. a Young specimen (Buyck 16.087) showing the typical dark violet pileus and short, fleshy, yellowish stipe as illustrated (Heinemann 1959) for the holotype. b Mature rusty brown discolored form (Buyck 16.080). c Mature yellow discolored form (Buyck 16.096). d Detail of the hairy pileus covering in young specimens (Buyck 16.087). e Epitype (Buyck 16.064). f typical, amethysteus-like form (Buyck 10.063). g detail of a mature specimen (Buyck 16.127) with a strongly veined hymenophore. Photos B. Buyck. Scale bars: a–g = 1 cm in particular because both share a strongly veined-anastomosing hymenophore that is off-white when young. However, C. goossensiae forms much more robust, fleshy fruiting bodies that lack the long and slender stipe and the very thin pileus typical of the latter species. Fully grown specimens of C. goossensiae are also more variable in color compared to C. miniatescens as the latter never develops the brown to olive tinges observed in several collections of the former (see Fig. 144b, c, e). Some collections of C. goossensiae developed a typical orange tinge in its upper part of the stipe, something that appeared to be a good feature for the correct identification of specimens that lost 123 202 Fungal Diversity (2019) 96:1–242 Fig. 145 Cantharellus goossensiae (epitype). Microscopic features. a Spores. b Basidia and basidiola. c Hyphal extremities of the pileus surface. Drawings B. Buyck. Scale bar = 10 lm, but only 5 lm for spores their violet pileus color. In addition, C. goossensiae generally lacks the dispersed remains of the red fibrillose tomentum that are visible near the cap centre or near the very cap margin of mature or older C. miniatescens (see for illustrations Buyck et al. (2016a)). The latter species also has narrower spores and more slender hyphal extremities composed of more elongated cells at the pileus surface. 123 Cantharellus brunneopallidus Buyck, Randrianjohany & V. Hofst., sp. nov. MycoBank number: MB829484; Facesoffungi number: FoF05997; Figs. 147, 148a–c Etymology: Referring to the brownish color when young, fading to yellowish cream, isabelline or dirty whitish with age. Holotype: MADAGASCAR. Tampolo, S 17.28724 – E49.40868, growing on sandy soil between grass and Fungal Diversity (2019) 96:1–242 amongst dead leaves under Intsia, 4 July 2011, Buyck, Hofstetter & Randrianjohany leg., Buyck 1053/BB11.105 (PC0085584, holotype). Fruiting bodies rather small, less than 30 mm high, in groups of several to many dispersed to aggregated individuals. Pileus less than 30 mm diam., regular in outline, more rarely lobed or wavy–undulate, first convex, then becoming progressively more depressed in the centre, finally often infundibuliform; margin strongly inrolled, plane or slightly wavy, the marginal zone of the cap remaining often downturned in age; surface finely fibrillose, slimy when wet, when young often unevenly coloured in patches of yellowish brown to reddish brown or even dark brown (5CD6–7,7DEF6–8), later sometimes locally retaining these colors, elsewhere gradually paler, discoloring to pale cream or isabelline between pale brownish squamae. Hymenophore decurrent, rather low and mostly ca 1 mm high, quite abruptly delimited from the sterile stipe surface, composed of crowded (25–35/cm), frequently forking and thin gill folds, whitish and with concolorous, even edges, without anastomosing–veined pattern in between gills or on gill sides. Stipe shorter than the cap diam., 9–15(19) 9 3–5(7) mm, subcylindrical or slightly narrowing downward, slimy when wet, particularly in lower half also covered with similar, fibrillose–erected to appressed, pale brownish squamulae as the pileus surface, off–white, but soon yellowing with age below, sometimes brownish in lower half when strongly squamulose, not hollowing. Context white, rather firm, relatively thick (2–3 mm just outside the pileus centre), strongly bruising chrome yellow when handled on the whole surface of the fruiting bodies or in lower tissues from insect damage, but not entirely becoming yellow. Taste mild. Smell weak. Spore print not obtained. Spores ellipsoid, (6.2)6.4–6.68–6.9(7.0) 9 (3.9)4.2– 4.47–4.8(5.0) lm, Q = (1.36)1.41–1.50–1.59(1.70), smooth, hyaline. Basidia small and short, mostly 35–42 9 6–7 lm, usually six–spored. Cystidia none. Subhymenium more pseudoparenchymatic than filamentous. Pileipellis composed of thin-walled to refringent, intertwining, septate and mostly slender generative hyphae, (4)6–9(12) lm diam., with many free endings and occasional ramifications; terminal cell mostly (30)40–65 lm long, rounded obtuse or slightly tapering at the apex, thinwalled to refringent, but very few clearly thick-walled, mostly quite regular in outline, sometimes locally undulating or more abruptly inflated. Clamp connections absent from all tissues. Additional material examined: MADAGASCAR. Tampolo, S 17.28724 – E49.40868, growing on sandy soil between grass and amongst dead leaves under Intsia, 6 July 2018, 1054/Buyck 11.116 (PC0085585, paratypus). 203 Notes: The description is based on the holotype. The paratype has similar features, including a near identical spore size of (6.0)6.3–6.66–7.0(7.3) 9 (3.9)4.1–4.49– 4.7(4.8) lm, Q = (1.35)1.42–1.52–1.63(1.70). This species is strongly reminiscent of the tropical African C. densifolius (sensu lato) because of the equally crowded gill folds and overall color, and it is thus not surprising that it was originally mistaken for that species when we collected it in the field. Nevertheless, the very strong yellowing when bruised is more typical of the Malagasy C. albidolutescens Buyck, Eyssart. & V. Hofst., but the latter has less crowded gill folds (see Buyck et al. (2015)). Our phylogeny now shows that it is clearly an independent species. As we never have found C. brunneopallidus in other habitats or with other host trees during the many years of collecting in Madagascar, it is possible that it might be a strict associate on the island of the ectomycorrhizal tree genus Intsia Thouars (Fabaceae), a very small tree genus with an indopacific distribution (Asamoah et al. 2012), but more sites with Intsia need to be studied in Madagascar to confirm this. Intsia bijuga (Colebr.) Kuntze, the host species on our collecting site, is known to have survived in a few remaining remnants of low elevation humid forests along Madagascar’s east coast. This tree was always heavily exploited for its high quality wood. It is listed for many years already on the IUCN Red List of Threatened Species and is believed to risk worldwide extinction within the next 20 years without appropriate action to stop illegal logging (see Intsia bijuga page on Wikipedia for more details). Cantharellus griseotinctus Buyck, Randrianjohany & V. Hofst. sp. nov. MycoBank number: MB829483; Facesoffungi number: FoF05998; Figs. 148d–f, 149 Etymology: griseo- and –tinctus, referring to the greyish tinges of the pileus surface. Holotype: PC0085579 Diagnosis: differs from C. albolutescens in the less crowded and better developed, less forked gill folds with distinct rib-like anastomoses on the gill sides, as well as in the much shorter terminal cells of the hyphal extremities in the pileipellis. Fruiting bodies medium-sized, growing single or in small groups. Pileus relatively fleshy and firm, up to 45 mm diam., first convex, than plano–convex or becoming slightly depressed in the centre, surface a pale brown to brownish gray (6–7DE4–7) or pale gray to grayish cream toward the margin, from centre to margin gradually more areolate–fissured, pileus centre distinctly warty–scurfy, fading to hardly visible squamae to almost smooth toward the margin, rarely developing strongly strigose–squamose rims in concentrical arrangement. Hymenophore decurrent, composed of well–developed, relatively thick gill folds, up 123 204 123 Fungal Diversity (2019) 96:1–242 Fungal Diversity (2019) 96:1–242 b Fig. 146 Most likely phylogram (-ln = 12014.27555) inferred from combined analyses of nucLSU and TEF1-a for 106 cantharelloid taxa, using Craterellus tubaeformis as the outgroup taxon. The combined gene sequence analyses comprise 1840 characters after exclusion of ambiguously aligned regions (1211 characters for nucLSU, 629 characters for TEF1-a). Phylogenetic analyses were all conducted in PhyML (Guindon and Gascuel, 2003) under GTR evolutionary model with the number of substitutions categories, the proportion of invariable sites and the gamma shape parameters estimated during the searches. No significant conflict was detected between nucLSU and TEF1-a based on comparison of bootstrap values recovered from bootstrap analyses of individual gene (200 bootstrap replicates with conflict assumed when two different relationships, one being monophyletic and the other being non-monophyletic, for the same set of taxa were both supported with significant bootstrap values equal or greater than 70%; Mason-Gamer and Kellog 1996). Searches for the most likely tree included three independent runs. Model parameters for the most likely tree were as follows: proportion of invariable sites = 0.412; gamma shape parameter = 0.409; A = 0.26337, C = 0.22919, G = 0.27754, T = 0.22990 for base frequencies; A– C = 1.09799, A–G = 5.55043, A–T = 1.56560, C–G = 1.10930, C– T = 10.70297, G–T = 1.00000 for substitution rates. The newly generated sequences are indicated in bold and the taxa discussed in bold blue. GenBank accessions for nucLSU/TEF1-a follow taxon names. Branches that received significant support (equal or greater than 70%) based on 500 BS replicates are in bold and BS values indicated along the branches to 4 mm high, unequal and shorter gills of variable length present, radially undulate –wavy due to rib–like veins running down the gill sides, off–white to pale cream with even smooth edges, furcations frequent only near the pileus margin. Stipe stout, shorter than the pileus diam., less than 20 mm long, 4–7 mm diam., surface white, minutely warty – fibrillose, not hollowing. Context relatively firm, white, but near the surface strongly bruising yellow when handled. Taste mild. Smell weak. Spore print not obtained. Spores rather small, (6.0)6.3–6.64–7.0(7.5) 9 (3.7) 4.0–4.43–4.8(5.2) lm, Q = (1.30)1.39–1.51–1.62(1.72), ellipsoid, smooth, with a small apiculus. Basidia clavulate, small to almost medium–sized, 35–50(60) 9 7–8 lm, mostly six–spored; sterigmata robust. Cystidia none. Subhymenium more pseudoparenchymatic than filamentous. Pileipellis composed of ramifying, intermingled, strongly thick–walled hyphal extremities, that are densely septate (with [ 10 consecutive thick–walled cells) but easily fragmenting when making preparations, sparsely branching, measuring mostly (4)6–8(10) lm diam.; cells becoming gradually shorter toward the apex, with the terminal cell rarely longer than 30 lm, tapering and more or less conical, or also sometimes more ellipsoid and inflated. Clamp connections absent from all tissues. Materials examined: MADAGASCAR. East Coast, Analalava, 7 km West of Mahovelona (Foulpointe), 1742S-4927E, in low altitude humid forest, in small 205 groups among herbs on muddy, stony, steep slope, Buyck 1048/BB 11.054 (PC0085579, holotype !). Notes: This species strongly resembles C. albidolutescens, (see Buyck et al. (2015)) but it differs from it because of the less crowded and better developed, less forked gill folds with distinct rib-like anastomoses on the gill sides. The more spaced gills also immediately exclude C. densifolius and look-alikes as a possible identification. It was collected on the outer limit of one of the few remnants of low altitude humid forests on the east coast. This forest remnant near Analalava, not far from Foulpointe, is known for its high biodiversity and conservation value, which is well documented in particular for the many rare palm species that are growing there (Rakotoarinivo et al. 2010). Hymenochaetales Oberw. Notes: The order was set up by Frey et al. (1977), based on Hymenochaetaceae Donk, the most species in the order have xanthochroic reaction. But recent molecular analyses showed that more families without xanthochroic reaction were included in the order (Miettinen and Larsson 2011). Hymenochaetaceae Donk, Bull. bot. Gdns Buitenz. Notes: The family Hymenochaetaceae was introduced by Donk (1948), based on the type genus Hymenochaete Lév. It is the most important family in Hymenochaetales, because it has the major species in the order, many species in the family are medicinal fungi and some are forest pathogens (Dai et al. 2007, 2009). The xanthochroic reaction is the most important characteristic of the family. Currently around 20 genera and 500 species are accepted in the family (Larsen and Cobb-Poulle 1990; Léger 1998; Parmasto 2005; Kirk et al. 2008; Dai 2010; Cui et al. 2011; He and Dai 2012; Ji et al. 2017, 2018; Zhou et al. 2016a, b, c). Fomitiporia Murrill, N. Amer. Fl. Notes: Fomitiporia was included in Phellinus Quél. sensu lato previously, but is characterized by subglobose to globose, hyaline, thick-walled, strongly dextrinoid and cyanophilous basidiospores, and accepted as an independent genus (Fiasson & Niemelä 1984; Dai 2010; Amalfi & Decock 2013; Chen & Cui 2017). Fomitiporia carpinea X.H. Ji, X.M. Tian & Y.C. Dai, sp. nov. Index Fungorum number: IF555462; Facesoffungi number: FoF04971; Fig. 150 Etymology: Referring to the species growing on Carpinus. Holotype: BJFC 025552. Basidiocarp perennial, resupinate, inseparable, without odour or taste when fresh, consistency woody hard, light in weight when dry, up to 10 cm long, 4 cm wide and 7 mm thick at centre, becoming more or less cushion-shaped with 123 206 Fungal Diversity (2019) 96:1–242 Fig. 147 Cantharellus brunneopallidus (PC0085584, holotype). Microscopic features. a Spores. b basidia and basidiola. c Hyphal extremities at pileus surface. Drawings B. Buyck. Scale bar = 10 lm, but only 5 for spores age; margin receding, brown, up to 2 mm wide. Pore surface dark brown to bay when fresh, becoming yellowish brown to umber up on drying, distinctly shining; pores circular, 4–6 per mm; dissepiments thin and entire. Subiculum brown, up to 2 mm thick. Tubes concolorous with pore surface, stratified, up to 5 mm long. Hyphal system dimitic in all parts; generative hyphae simple septate, skeletal hyphae negative in Melzer’s reagent, acyanophilous in Cotton Blue; tissue darkening but otherwise unchanged in KOH. Subicular generative hyphae 123 infrequent, hyaline to pale yellowish brown, thin- to slightly thick-walled, occasionally branched, simple septate, 3–4 lm in diam.; skeletal hyphae yellowish brown, thick-walled, with a wide lumen, unbranched, regularly arranged, 3–5 lm in diam. Tramal generative hyphae hyaline to pale yellow brown, thin- to slightly thick-walled, occasionally branched, frequently simple septate, 2.5–3.5 lm in diam.; skeletal hyphae yellowish brown, thick-walled, with a narrow lumen, rarely branched, interwoven, 3–4.5 lm in diam. Hymenial setae absent; Fungal Diversity (2019) 96:1–242 207 Fig. 148 Cantharellus brunneopallidus (PC0085584, holotype). a Freshly collected specimens. b Young specimen showing the dark pileus. c holotype overview. d–f Cantharellus griseotinctus (PC0085579, holotype). d showing color change after one day with respect to the freshly collected specimens depicted in e–f (photos B. Buyck) Scale bars: 1 cm cystidioles present, ventricose, hyaline, thin-walled, 10–18 9 3–5 lm; basidia short clavate to capitate, with a basal simple septum and four sterigmata, 12–16 9 8.5–11 lm; basidioles in shape similar to basidia, but slightly smaller, big irregular or rhombic crystals present among trama. Basidiospores globose, hyaline, thick-walled, smooth, dextrinoid in Melzer’s reagent, cyanophilous in Cotton Blue, (6.1–)6.5–7.8(–8) 9 (5.5–)6–7.5(–7.8) lm, L = 7.06 lm, W = 6.7 lm, Q = 1.05 (n = 30/1). Material examined: CHINA, Gansu Province, Pingliang, Kongtongshan Forest Park, on stump of Carpinus, 2 Sept 2017, Dai 18023 (BJFC 025552, holotype). GenBank numbers: ITS: MH930812, LSU: MH930810. Notes: Fomitiporia carpinea is similar to F. punctata (P. Karst.) Murrill in sharing perennial, resupinate, cushion- 123 208 Fungal Diversity (2019) 96:1–242 Fig. 149 Cantharellus griseotinctus (PC0085579, holotype). Microscopic features. a Spores. b basidia and basidiola. c Hyphal extremities at pileus surface. Drawings B. Buyck. Scale bars = 10 lm, but only 5 for spores shaped basidiocarps, the approximately the same size basidiospores, and occurring in temperate forests (Dai 2010, 2012). However, the latter species has smaller pores (6–8 per mm), interwoven hyphae in subiculum, lacks cystidioles, while F. carpinea has bigger pore (4–6 per mm), hyphae regularly arranged in subiculum, the presence of cystidioles. Phylogentically F. carpinea formed a distinct terminal lineage (Fig. 152), and is distant from F. punctata. 123 Fomitiporia lagerstroemiae X.H. Ji, X.M. Tian & Y.C. Dai, sp. nov. Index Fungorum number: IF555461; Facesoffungi number: FoF04972; Fig. 151 Etymology: Referring to the species growing on Lagerstroemiae. Holotype: BJFC 025858 Basidiocarp perennial, resupinate, inseparable, without odour or taste when fresh, consistency woody hard, light in weight when dry, up to 5 cm long, 3 cm wide and 8 mm Fungal Diversity (2019) 96:1–242 209 Fig. 150 Microscopic structures of Fomitiporia carpinea (BJFC 025552, holotype). a Basidiospores. b Basidia. c basidioles. d Cystidioles. e Hyphae from trama. f Hyphae from subiculum thick at centre, becoming more or less cushion-shaped with age; margin receding, dark brown, up to 2 mm wide. Pore surface snuff brown when dry, distinctly shining; pores circular, 7–9 per mm; dissepiments thin and entire. Subiculum brown, very narrow to almost lacking. Tubes concolorous with pore surface, stratified, up to 8 mm long. 123 210 Fungal Diversity (2019) 96:1–242 Fig. 151 Microscopic structures of Fomitiporia lagerstroemiae (BJFC 025858, holotype). a Basidiospores. b Basidia. c Basidioles. d Hymenial setae. e Hyphae from subiculum Hyphal system dimitic in all parts; generative hyphae simple septate, skeletal hyphae negative in Melzer’s reagent, acyanophilous in Cotton Blue; tissue darkening but otherwise unchanged in KOH. Subicular generative hyphae infrequent, hyaline to pale yellowish brown, thinto slightly thick-walled, occasionally branched, simple septate, 2.5–4 lm in diam.; skeletal hyphae yellowish brown, thick-walled, with a wide lumen, unbranched, loosely interwoven, 3–4.5 lm in diam.Tramal generative 123 hyphae hyaline to pale yellow brown, thin- to slightly thick-walled, occasionally branched, frequently simple septate, 2.5–3 lm in diam.; skeletal hyphae yellowish brown, thick-walled, with a narrow to medium size lumen, rarely branched, interwoven, 3–4 lm in diam. Hymenial setae frequent, ventricose, thick-walled, dark brown, 15–22 9 5–7 lm; cystidioles absent, basidia barrelshaped, with a basal simple septum and four sterigmata, 9.5–11.5 9 6.5–9 lm; basidioles in shape similar to Fungal Diversity (2019) 96:1–242 211 basidia, but slightly smaller. Basidiospores subglobose, hyaline, thick-walled, smooth, dextrinoid in Melzer’s reagent, cyanophilous in Cotton Blue, (4.8–)5–6(–6.2) 9 (4–)4.5–5.5(–6) lm, L = 5.86 lm, W = 5.13 lm, Q = 1.14 (n = 30/1). Material examined: VIETNAM, Dong Nai Province, Dinh Quan District, Thac Mai Preservation Park, on stump of Lagerstroemia, 14 Oct 2017, Dai 18335 (BJFC 025858, holotype). GenBank numbers: ITS: MH930812, LSU: MH930810. Notes: Fomitiporia lagerstroemiae is similar to Fomitiporia hainaniana B.K. Cui & Hong Chen in sharing perennial, resupinate, basidiocarps, the approximately the same size pores, the presence of hymenial setae, interwoven tramal hyphae and occurring in tropical forests (Chen and Cui 2017). However, Fomitiporia hainaniana has smaller basidiospores (4–5 9 3.8–4.4 lm), and grows on angiosperm wood rather than Lagerstroemia (Chen and Cui 2017). The phylogentic analysis of ITS and LSU sequence (Fig. 152) shows that Fomitiporia lagerstroemiae is closely related to Fomitiporia bannaensis Y.C. Dai, and both species has perennial, resupinate, basidiocarps, the approximately the same size pores and hymenial setae. However, Fomitiporia bannaensis has smaller Fig. 152 Phylogenetic position of Fomitiporia carpinea (holotype) and Fomitiporia lagerstroemiae (holotype) inferred from the ITS and LSU sequences. Bootstrap support values for ML and MP greater than 50% and Bayesian posterior probabilities greater than 0.95 are given near nodes respectively. The tree is rooted with Phellinus uncisetus (MUCL 46231) and P. uncisetus (MUCL 47061). The new isolates are in bold and black 123 212 basidiospores (4.2–5.2 9 3.8–4.9 lm, L = 4.68 lm, W = 4.24 lm), the presence of cystidioles (Dai 2010). Polyporales Gäum. Notes: Polyporales is a large group of Agaricomycetes with about 1800 described species (Kirk et al. 2008; Justo et al. 2017). This order includes species with a wide variety of basidiomata (pileate, stipitate, resupinate or multiple flabelliform lobes) and hymenophore (poroid, hydnoid, lamellate, merulioid or smooth) shapes (Binder et al. 2013; Hibbett et al. 2014). Most of the species are saprotrophic wood-decay fungi, while few are plant pathogens. Polyporales has been the focus of many phylogenetic studies which have confirmed it as a strongly supported monophyletic clade of Agaricomycetes (Binder et al. 2013; Hibbett et al. 2014; Justo et al. 2017). Polyporaceae Fr. ex Corda Notes: Polyporaceae (Polyporales) was proposed by Fries (1838) with Polyporus P. Micheli as the type genus. The family currently includes about 45 genera (Justo et al. 2017). Phylogenetic analyses have shown that Polyporaceae belongs to the ‘‘core polyporoid’’ within Polyporales (Binder et al. 2013). Species in the family have poroid, irregular or lamellate, occasionally corticioid hymenophore. The hyphal system is rarely monomitic, mostly dimitic or trimitic, and cystidia are generally lacking. The species are saprobes and cause white rot (Cannon and Kirk 2007; Justo et al. 2017). Grammothele Berk. & M.A. Curtis Notes: Grammothele, typified by G. lineata Berk. & M.A. Curtis (Berkeley and Curtis 1869), is a genus with about 22 species (www.indexfungorum.org), with thin and Fungal Diversity (2019) 96:1–242 strongly adherent basidiomata, and shallow pores (Ryvarden 2015). The hyphal system is dimitic to trimitic, with clamped generative hyphae, skeletal hyphae thick-walled to solid, hyaline, darkening with age, dextrinoid or not. Dendrohyphidia and others sterile elements are absent or present and the basidiospores are ellipsoid to cylindrical, thin-walled, smooth and non-amyloid. This tropical genus occurs on mono- and dicotyledons (Karasinsk 2015; Ryvarden 2015; Wu et al. 2016). Grammothele aurantiaca A.M.S. Soares, sp. nov. Index Fungorum number: IF555479; Facesoffungi number: FoF05974; Fig. 153 Etymology: aurantiacus L. = orange, referring to the colour of the basidiomata. Holotype: MG 232015 Basidiomata annual, resupinate, strongly adnate, hard and brittle when dried, up to 8 cm long, 4.5 cm wide, 1 mm thick; pore surface salmon to orange (5C5), becoming dark orange with the age (5B8); margin narrow to almost lacking, rust orange (5C4). Pores shallow, almost invisible to the naked eye, round to angular, 6–8 per mm, dissepiments thin, tubes concolorous, up to 1 mm deep. Basidiospores scarce, subglobose, hyaline, thin-walled, with tiny or without apiculus, smooth, IKI-, 4–5 lm wide (L = 4.66 lm, W = 4.33 lm, Q = 1.07, n = 10/2). Basidia not seen. Hyphal system dimitic, generative hyphae thickwalled, up to 3 lm wide, without clamps, IKI-; skeletal hyphae thick-walled 4–5 lm, tightly interwoven, cherry red in KOH, IKI-. Dendrohyphidia absent. Cystidia and others sterile elements absents. Material examined: BRAZIL, Amapá: Serra do Navio, Serra dos Veados, on dead wood, 00 550 39,9600 N and 51 Fig. 153 Grammothele aurantiaca (MG 232015, holotype). a Basidioma. b Pores. c Hyphae from the sub-hymenium in KOH d basidiospores. Photos: A.M.S. Soares; Drawing: R.L. Alvarenga. Scale bars: a = 1 cm, b, c = 10 lm, d = 5 lm 123 Fungal Diversity (2019) 96:1–242 590 20,7900 W, October 2014, W. Xavier, WX 2014-115, (MG 232015, holotype; isotype in URM). Additional specimens examined: Grammothele aurantiaca – BRAZIL, Amapá: Serra do Navio, Serra dos Veados, on dead wood, February 2015, W. Xavier, WX 2015-88 (MG 226148). Porogramme lateritia – COSTA RICA, Alajuela: Bijagua, Parque Nacional Volcan A renal La Fortuna de San Carlos, July 2001, L. Ryvarden, 43 (O F18637). DOMINICAN REPUBLIC: November 1892, Elliot (O F903299). FRANCE, Guadeloupe: Camp Jacob, sur un tronc d’un Synplocus maritimiersii, Duss 592 (O F505994, holotype); Basse terre: October 1976, A. David, 2189 LY-AD (O F505992). VENEZUELA, Arugua: dead hardwood, February 2006, L. Ryvarden (O F506199). GenBank numbers: LSU: MH844886, ITS: MH842137. Notes: Grammothele aurantiaca is characterized by the salmon to dark orange basidiomata, dimitic hyphal system, generative hyphae simple-septate, and skeletal hyphae cherry red in KOH. Porogramme lateritia (Pat.) Pat. has similar colour, however the pores are smaller [10–15(–20 per mm)], the hyphal system is monomitic with clamps and the basidiomata become greyish to reddish when dried (Lowe 1964; Ryvarden 1983). Besides, in the phylogenetic tree (Fig. 156), G. aurantiaca clustered in the Grammothele clade with good support (1.00/89%) and is distantly 213 related to the Porogramme clade, with P. albocincta (Cooke & Massee) Gibertoni (1.00/95%) (Fig. 156) as the type species of the genus. Grammothele micropora A.M.S. Soares & W.K.S. Xavier sp. nov. Index Fungorum number: IF555480; Facesoffungi number: FoF05975; Fig. 154 Etymology: mikrós Gr. = small ? porus L. = pore, referring to the tiny pores. Holotype: MG 232012 Basidiomata annual, resupinate, effuse, strongly adnate, hard and brittle when fresh and when dried, up to 15 cm long, 8 cm wide, 1 mm thick; pore surface black when fresh, dark-bluish when dried (1F1); margin narrow, concolorous with the pore surface becoming slightly whitish with the age. Pores shallow, invisible to the naked eye, irregular, round, sometimes angular to hexagonal, 25–30 per mm, dissepiments thin, tubes concolorous, up to 1 mm deep. Basidiospores subglobose, hyaline, thick-walled, smooth, with tiny or without apiculus, IKI-, (3.0–) 3.5–4(– 4.5) 9 3–3.5 (–4) lm (L = 3.81 lm, W = 3.14 lm, Q = 1.21, n = 15/1). Basidia narrowly clavate, 7–10 9 3–5 lm, with four sterigmata, 2.5–3 lm long. Hyphal system dimitic, generative hyphae, thin-walled, Fig. 154 Grammothele micropora (MG 232012, holotype). a Basidiomata. b Pores. c Hyphae from the sub-hymenium. d1 Cystidioles. d2 Basidia. e Basidiospores. Photos: A.M.S. Soares; Drawings: R.L. Alvarenga. Scale bars: a = 1 cm, b, c = 10 lm, d, e = 5 lm 123 214 hyaline, clamped, 2–3 lm wide, IKI-; skeletal hyphae dominating, thick-walled, brown to black in KOH, 5–7 lm wide IKI-. Dendrohyphidia absent. Cystidia absent, but fusoid cystidioles present, 6–9 9 4–5 lm. Material examined: BRAZIL, Amapá: Serra do Navio, 0 550 9.9500 N and 51 590 19.800 W, on dead wood, October 2014. W. Xavier, WX 2014-116 (MG 232012, holotype; isotype in URM). Additional specimens examined: BRAZIL, Amapá: Porto Grande, Floresta Nacional do Amapá, July 2009, H. Sotão, H2009-256A (MG 232558). BRAZIL, Amapá: Porto Grande, Floresta Nacional do Amapá, September 2011, A.M.S. Soares, M146 (MG 232692). GenBank numbers: ITS: MH842144, LSU: MH842144. Notes: the dark bluish to black basidiomata, the small pores and basidiospores, among the smallest in the genus, are characteristic of this species. Initially, G. micropora was identified as Porogramme albocincta (Cooke & Massee) Gibertoni because of its dark bluish grey basidiomata and very small pores (8–20 per mm). Nevertheless, G. micropora has even smaller pores and dimitic hyphal system, whereas P. albocincta is monomitic and its basidiospores are longer [4–6 (6.5) 9 3–3.5 lm] (Ryvarden 1979). Grammothele boliviana Karasiński has similar colour (grey to bluish grey), however the basidiospores are broadly ellipsoid to ovate and larger [8.4–10.8(–11.5) 9 6–7.2(– 7.5) lm]. Differently from the new species, G. boliviana has hyphidia and dendrohyphidia and is reported on palms Fungal Diversity (2019) 96:1–242 (Karasinsk 2015). Additionally, in the phylogenetic tree, G. micropora clustered within the Grammothele clade (1.00/ 76%) and grouped with low to moderate support with G. brasiliensis Ryvarden (0.95/70%) (Fig. 156). The type species of the G. brasiliensis is from Brazil’s southeast (state of São Paulo) and it is characterized by the dark grey basidiomata, pores 5–6 per mm and cylindrical basidiospores (5–6 9 2.5–3 lm) (Ryvarden 2015). Grammothele brasiliensis Ryvarden [as ‘brasilensis’], Syn. Fung. (Oslo) 33: 38 (2015) Facesoffungi number: FoF05976; Fig. 155 Description: Ryvarden (2015). Material examined: BRAZIL, Amapá: Serra do Navio, on dead wood, October 2014, W. Xavier, WX2014-87 (MG232006), WX2014-132 (MG232007), WX2014-100 (MG232008), WX2014-28 (MG232009), WX2014-101 (MG232010); WX2014-148 (MG232011). Grammothele subargentea – BRAZIL, Amapá: Serra do Navio, on dead wood, October 2014, W. Xavier, WX2014-26 (MG232013), WX2014-155 (MG232014). Grammothele lineata – BRAZIL, Amapá: Serra do Navio, on dead wood, October 2014, W. Xavier, WX2014-208 (MG232015), WX2014-40 (MG232016). GenBank numbers: ITS: MH844864, MH844550, MH844679, MH844866, MH844553, MH844563, MH842150, LSU: MH844863, MH844887, MH844583, MH844865, MH844553, MH842149. Fig. 155 Grammothele brasiliensis (MG 232007, new record). a Basidioma. b Pores. c Encrusted hyphae from the sub-hymenium. d Basidia. e Basidiospores. Photos: A.M.S. Soares; Drawings: R.L. Alvarenga. Scale bars: a = 2 cm, b = 0.2 mm, c = 10 lm, d, e = 5 lm 123 Fungal Diversity (2019) 96:1–242 Notes: This species is characterized by the dark grey basidiomata, round and entire pores (5–6 per mm), dimitic hyphal system, clamped generative hyphae, skeletal hyphae strongly agglutinated with numerous crystals and cylindrical basidiospores (5–6 9 2.5-3 lm) (Ryvarden 2015). Dendrohyphidia were easily identified on the studied material, which were not found by Ryvarden (2015). He, however, presumed their presence in the dissepiments. This species may be confused with Tinctoporellus epimiltinus (Berk. & Br.) Ryvarden, which has similar colour (bluish white and light beige), pores (angular to round, 7–9 per mm), and dimitic hyphal system. However, the ellipsoid to subglobose basidiospores [4–5 (5.5) 9 2.5–3 lm], absence of dendrohyphidia and cystidioles separate it from G. brasiliensis (Ryvarden 1979). Besides, G. brasiliensis did not group in the Tinctoporellus clade (1.00/95%), but formed a monophyletic lineage with strong support (1.00/ 100%) within the Grammothele clade (Fig. 156). This is the first time this species is sequenced, the second record of this species for Brazil and the first to the Brazilian Amazonia. The colour of G. subargentea (Speg.) Rajchenb. resembles G. brasiliensis, but the former has larger pores and longer basidiospores [(2–3 per mm; cylindrical (6–) 6.5–8.5 9 (–2.5) 3–4 lm)]. The type species of G. subargentea is from Paraguay, and the morphological characters of the specimens analyzed in this study are very similar to the original description (Ryvarden 2015). Although there is no sequence of the type locality, our phylogenetic analyses confirm this species in the Grammothele clade (0.91/69%). Another species registered in Brazil, G. lineata Berk. & M.A. Curtis, has also white to greyish basidiomata, later pinkish, and might be mistaken for G. brasiliensis; however, the former has trimitic hyphal system and a partly hydnoid surface with dots of numerous dark bundles of skeletal hyphae (Ryvarden 2015). In addition, the specimens collected in the Brazilian Amazonia sistered with G. lineata from China and Malaysia, and Grammothele sp. and G. denticulata Y.C. Dai & L.W. Zhou (1.00/99%) from China, in the Grammothele clade. Grammothele lineata was originally collected in Cuba and there is no sequence from the type locality. The new sequences generated from Brazilian material may represent G. lineata due the closeness to the type locality, since both are in the Neotropics. Thus, the Asian collections should be reevaluated. Russulales Kreisel ex P.M. Kirk, P.F. Cannon & J.C. David Notes: The concept of Russulales was gradually developed in morphological analyses for several groups of macro-fungi. Singer and Smith (1960) have noticed the close morphological relationship between agaricoid and 215 gasteroid homobasidiomycetes taxa. The amyloid basidiospore ornamentations and gloeoplerous hyphae system reminded the Dutch mycologist Marinus Anton Donk of some families (e.g. Hericiaceae) may be members of an unknown order (Donk 1971). Oberwinkler (1977) regarded this order as Russulales which include agaricoid, discoid, fused-reflexed, gasteroid, hydnoid, and resupinate taxa. The order Russulales was officially acknowledged in the Dictionary of Fungi (Kirk et al. 2001; Miller et al. 2006). The independence of this order has been supported by some recent phylogenetic analyses (Hibbett et al. 2007; Zhao et al. 2017). Russulaceae Lotsy Notes: The family Russulaceae was established by Dutch botanist Johannes Paulus Lotsy to accommodate species with granular flesh, thick gills, spiny spores, and milky hyphae and rounded cells (sphaerocytes). This family includes gasteroid and agaricoid genera, Lactarius Pers., Lactifluus (Pers.) Roussel, Multifurca Buyck & V. Hofst. and Russula Pers., together with the crust-like genera Boidinia Stalpers & Hjortstam, Gloeopeniophorella Rick, and Pseudoxenasma K.H. Larss. & Hjortstam (Lotsy 1907; Larsson and Larsson 2003; Miller et al. 2006; Larsson 2007; Buyck et al. 2008, 2010; Verbeken and Nuytinck 2013). The family comprises ectomycorrhizal (Tedersoo and Nara 2010), arbutoid mycorrhizal (Smith and Read 2008), monotropoid mycorrhizal (Bidartondo 2005), orchid mycorrhizal (Dearnaley 2007), and saprotrophic, wood-degrading species (Larsson and Larsson 2003). Russula Pers. Notes: The genus Russula was introduced by Persoon (1796) with the type species R. emetica (Schaeff.) Pers. This genus is characterized by fairly large basidiocarps, brightly colored pileus, white to dark yellow basidiospore print, attached lamellae, absence of volva and milky latex, amyloid basidiospore ornamentations, and numerous spherocysts in context (Sarnari 1998). Over 800 species are accommodated in this genus (Li 2014; Li et al. 2015, 2018c). Most recent new Russula species were described based on specimens from Asia (Li et al. 2016, 2018a, b; Sang et al. 2016; Das et al. 2017a, b, 2018b; Hyde et al. 2016, 2017b; Jiang et al. 2017; Zhang et al. 2017). In this paper, we introduce a new species in the genus Russula which was collected in coniferous and broad-leaved intermixed forests from Guangxi of South China. Russula prasina G.J. Li & R.L. Zhao, sp. nov. Index Fungorum number: IF570591; Facesoffungi number: FoF05086; Fig. 157 123 216 Fig. 156 Maximum likelihood consensus phylogenetic tree of the concatenated ITS1, 5.8S, ITS2 and partial 28S rDNA sequences. Bootstrap supporting values (1000 replicates) and posterior probabilities (PP) from Bayesian analysis to each node are shown from left to 123 Fungal Diversity (2019) 96:1–242 right. Only bootstrap values above 50% and BYPP above 0.75 are provided. The newly generated sequences are indicated in bold and black. The tree was rooted with Funalia trogii (XSD37) Fungal Diversity (2019) 96:1–242 217 Fig. 157 Russula prasina (HMAS 281232, holotype). a, b Basidiocarps. c Basidiospores. d Basidia. e Subpellis. f Epipellis. Scale bars: a, b = 5 cm, f = 20 lm, d, e = 10 lm Etymology: Refers to the grass green tinged pileus of this species. Holotype: HMAS 281232 Basidiocarps small to large sized. Pileus 43–130 mm in diam., first hemispheric, then plane to depressed at centre when mature, slightly striate 25–30 mm towards the edge when mature, not cracked, slightly viscid when wet, dull, peeling 1/5–1/4 from the edge, grass green tinge of Winter Green (XVIII330 i), Motmot Green (XX430 i) to Courge Green (XVII250 i) at centre, pale yellowish green tinge of Oil Yellow (V27 k), Javel Green (V27i) to Calliste Green (VI31i) towards margin. Lamellae adnate, 3–5 mm in height, 8–13 pieces per centimeter at edge, rarely forked near stipe, often interveined, pale ocher tinge of Ochraceous-Buff (XV150 b), unchanging when bruised, lamellulae absent. Stipes central to subcentral, 4.1–7.7 9 2.4–3.6 cm, subcylindrical to cylindrical, surface dry, rugulose longitudinally, dull, without annulus, slightly attenuate downward, White (LIII) at upper part, ochraceous yellow tinge of Light Ochraceous Salmon (XV130 b) to Salmon Buff (XIV110 d) towards the base, stuffed at first, becoming hollow when old. Context up to 3–5 mm at 123 218 Fungal Diversity (2019) 96:1–242 Fig. 158 Phylogram generated from Bayesian analysis of ITS sequence data of Russula subgenus Heterophyllidia. Related sequences were obtained from Miller and Buyck (2002), Abarenkov et al. (2010), Li et al. (2013), Guo et al. (2014), Dutta et al. (2015), Zhao et al. (2015) and Zhang et al. (2017). Russula aurea and R. paludosa are used as the outgroup taxa. Bootstrap (BS) support values for ML equal to or greater than 65% and Bayesian posterior probability (BYPP) values equal to or greater than 0.90 are given above the nodes respectively as (BS/BYPP). Phylogenetic topology of the ML analysis was similar to that of the BI. The type specimens are in bold and black. The newly generated sequences are in blue centre of pileus, White (LIII) first, unchanging when bruised, fragile, no distinct odor; taste mild. Spore print Ocher (Romagnesi III b–c). Basidiospores [300/3/3] (5.9–) 6.2–7.3 (–7.6) 9 (5–) 5.4–6.5 (–7) lm, Q = 1.01–1.21 (–1.25) (Q = 1.12 ± 0.07), hyaline, globose to subglobose, rarely broadly ellipsoid; ornamentation composed of amyloid obtuse warts and short crests that are mostly interconnected as short crests and ridges, forming a complete network, warts 0.3–0.6 lm high; suprahilar area nonamyloid to rarely weakly amyloid. Basidia 35–53 9 8–11 lm, projecting 10–20 lm beyond hymenium, four-spored, sterigmata 4–6 lm long, hyaline, unchanging in KOH, subclavate to clavate, rarely cylindrical. Pleurocystidia and cheilocystidia not observed. Subhymenium cellular layer 15–30 lm thick composed of voluminous cells 10–20 lm in diam., hyaline, rarely pale yellowish in KOH. Pileipellis is composed of epipellis and subpellis; epipellis pseudoparenchymatous, 100–250 lm thick, composed of thinwalled, unbranched and inflated cells 10–20 lm wide, narrowing to 5–7 lm towards the terminal cells; pileocystidia absent; subpellis slightly gelified, a cutis 150–250 lm thick, composed of parallel to interweaved, rarely branched and septate, hyaline hyphae 2–4 lm wide. Stipitipellis cutis, upper layer composed of parallel filamentous hyphae 3–6 lm in diam., inner layer hyaline filamentous hyphae interweaved with sphaerocytes 15–25 lm in diam., some filamentous hyphae pale yellow in KOH; caulocystidia not observed. Trama composed of sphaerocytes 15–40 lm in diam. and scattered by filamentous and cystidioid hyphae. Clamp connections and laticiferous hyphae absent in all tissues. Material examined: CHINA, Guangxi Zhuang Autonomous Region, Baise City, Leye County, Huangjingdong National Forest Park, N 24740 E 106320 , elv. 378 m., 6 August 2017, Hui-Jun Wang GX20170580 (HMAS 281232, holotype); ibid. elv. 464 m., GX20170846 (HMAS 279805); ibid. elv. 395 m., GX20170937 (HMAS 279806). Habit and habitat: Scattered or single in coniferous and broad-leaved intermixed forests (dominated by e.g. Pinus 123 Fungal Diversity (2019) 96:1–242 219 Fig. 159 Rhizophydium koreanum (CNUFC-17CPW1-1, holotype). a Colony on PmTG agar. b Developing zoosporangium with branched rhizoids. c Resting spores. d Developing thallus with two rhizoidal axes (white arrows). e Uniflagellate zoospores (yellow arrows). Scale bars: b–e = 20 lm crassicorticea, P. yunnanensis var. tenuifolia, Quercus acutissima, and Q. variabilis) at 300–500 m altitude. Distribution: CHINA (Guangxi Zhuang Autonomous Region). Season. August. GenBank numbers: ITS: MH454351 (HMAS 281232), MH454352 (HMAS 279805), MH454353 (HMAS 279806). Notes: Russula prasina is a member of subgenus Heterophyllidia, section Virescentinae because it has greenish tinged pileus, small basidiospores, spherocysts in epipellis, and absence of pileocystidiain pileipellis (Sarnari 1998). Unlike the furfuraceous and areolate pileus surfaces of the other Virescentinae members, that of R. prasina is relatively smooth. The hyphal cells in subpellis and epipellis of R. prasina are reminiscent of those of R. parvovirescens Buyck, D. Mitch. & Parrent, in contrast, R. parvovirescens has pileocystidia in pileipellis and hymenium with pleurocystidia and cheilocystidia (Buyck et al. 2006). This new species may be confused with some green-capped Russula in China and adjacent region, e.g. R. atroaeruginea G.J. Li, Q. Zhao & H.A. Wen, R. dinghuensis J.B. Zhang & L.H. Qiu, R. nigrovirens Q. Zhao, Y.K. Li & J.F. Liang, R. sikkimensis K. Das, Atri & Buyck, R. virescens (Schaeff.) Fr., and R. viridirubrolimbata J.Z. Ying. These species can be distinguished from R. prasina by the following morphological characteristics: R. atroaeruginea differs in its radially yellowish-striped pileus, larger basidiospores (6.3–) 6.8–8.1 (–9.0) 9 (5.9–) 6.1–7.4 (–7.8) lm, and habitat of in alpine subtropical conifer forests dominated by Picea spp. (Li et al. 2013); R. dinghuensis has pileus surface cracking into small patches with rusty tinge, scattered lamellulae, white basidiospore sprint, and trichoderm epipellis (Zhang et al. 2017); R. nigrovirens can be differentiated by its longer basidia 45–75 9 9–14 lm, basidiospore ornamentation not forming a reticulum, and habitat of in alpine subtropical conifer forests dominated by Picea sp., Rhododendron sp., Sorbus sp. and Abies sp. (Zhao et al. 2015); R. sikkimensis has basidiospore ornamentation up to 0.9 lm, cutis to trichoderm pileipellis and habitat of in alpine subtropical conifer forests dominated by Abies spp. (Das et al. 2013); R. virescens differs in its white to pale cream spore print, 123 220 Fungal Diversity (2019) 96:1–242 Fig. 160 Phylogenetic tree of Rhizophydium koreanum (CNUFC17CPW1-1) and (CNUFC-17CPW1-2) and related species based on Maximum likelihood analysis of 28S. Sequence of Allomyces arbusculus was used as outgroup taxon. Numbers at the nodes indicate the bootstrap values (C 50%) from 1000 replications. The bar indicates the number of substitutions per position. The newly generated sequences are indicated in blue and ex-type strain in bold elliptical to subglobose basidiospores, and frequently elongated and tapered terminal cells in pileipellis (Sarnari 1998, Li 2014); R. viridirubrolimbata can be differentiated by its red tinge in pileus, and acrid context (Ying 1983). Phylogenetic analysis showed that our strains formed a distinct clade from known Russula species (Fig. 158). Thus Russula prasina is introduced as a new species based on morphological and phylogenetic analyses. Rhizophydium Schenk Notes: The genus Rhizophydium is the oldest genus within the order Rhizophydiales, proposed by Schenk (1858) and validated by Rabenhorst (1868). The species belonging to this genus are characterized by the formation of sphaerical zoosporangia with multiple, discharge pores, and branched rhizoids arising from the sporangium (Barr 1968; Letcher et al. 2006). Members of the genus occur in aquatic systems, primarily as parasites of algae, and from pollen, keratin, and soil as saprotrophs (Barr 1968; Longcore 1996; Letcher et al. 2004, 2006). In recent years, ultrastructural and molecular analyses have been used as a tool to delimit species in the genus, resulting in significant change in the taxonomy (Letcher et al. 2006; Seto et al. 2017). Many Rhizophydium species have been transferred to new genera (Letcher et al. 2006, 2015). During a study of Rhizophydiales in water samples in Gwangju, Korea, a new species of Rhizophydium was Chytridiomycota Arx Rhizophydiomycetes Tedersoo, Koljalg et al. Rhizophydiales Letcher Notes: Currently, 17 families and 25 genera are placed in order Rhizophydiales based on molecular phylogeny and zoospore ultrastructure (Letcher et al. 2015; Van den Wyngaert et al. 2017; Seto and Degawa 2018; Wijayawardene et al. 2018b). 123 Fungal Diversity (2019) 96:1–242 isolated and is described here, based on morphological characteristics and phylogenetic analyses. Rhizophydium koreanum Hyang B. Lee, S.J. Jeon, T.T. T. Nguyen, sp. nov. Index Fungorum number: IF554571; Facesoffungi number: FoF05792; Fig. 159 Etymology: koreanum, referring to the country from which the species was recorded. Holotype: CNUFC-17CPW1-1 On PmTG (peptonized milk, tryptone and glucose) agar: Zoosporangium sphaerical, with many closely spaced and highly branched rhizoids, measured (100–)107.5–120.5(– 131.5) lm diam. Rhizoidal system arising from a single point at the base of the sporangium. Zoospores abundantly produced, sphaerical, measured (3.0–)3.5(–4.0) lm diam., with a flagellum of (20–)23.5(–30) lm long. Resting spore sphaerical, measured (10–)12.5(–13.5) lm diam. Culture characteristics: Colonies reaching 8.5 mm diam. on PmTG at 25 C in 7 days, cream. Optimal growth was observed around 20 C. Material examined: REPUBLIC OF KOREA, from pond water collected at Chonnam National University Arboretum, Gwangju, Korea, 20 October 2017 (CNUFC17CPW1-1, preserved as glycerol stock at - 80 C in the Chonnam National University Fungal Collection; isotype in Culture Collection of Nakdonggang National Institute of Biological Resources [NNIBR], Sangju, Gyeongbuk Province). Notes: Based on phylogenetic analyses and morphological comparison, our isolate belongs to Rhizophydium. Rhizophydium koreanum formed a distinct clade from other species in the phylogenetic tree (Fig. 160). It differs from the closely related species R. globosum and R. brooksianum by forming larger sporangia. GenBank numbers: LSU: MH298649, MH298650. Acknowledgements Kevin D. Hyde thanks the Foreign Experts Bureau of Yunnan Province, Foreign Talents Program (2018; Grant No. YNZ2018002), Thailand Research grants entitled Biodiversity, phylogeny and role of fungal endophytes on above parts of Rhizophora apiculata and Nypa fruticans (Grant No. RSA5980068), the future of specialist fungi in a changing climate: baseline data for generalist and specialist fungi associated with ants, Rhododendron species and Dracaena species (Grant No. DBG6080013), Impact of climate change on fungal diversity and biogeography in the Greater Mekong Subregion (Grant No. RDG6130001). K.D. Hyde also thanks Chiang Mai University for the award of visiting Professor. We also thank to the director Jun-Bo Yang and Plant Germplasm and Genomics Center in Germplasm Bank of Wild Species, Kunming Institute of Botany for the molecular laboratory support. Dr. Shaun Pennycook is thanked for his essential nomenclatural review. Samantha C. Karunarathna thanks CAS President’s International Fellowship Initiative (PIFI) for funding his postdoctoral research (Grant No. 2018PC0006) and the National Science Foundation of China (NSFC, project code 31750110478). Department of Plant Medicine, National Chiayi University (NCYU) is thanked to provide facilities for DNA molecular experiment. Rajesh 221 Jeewon thanks Mae Fah Luang University for the award of a Visiting Scholar. Walter Rossi and Marco Leonardi wish to thank the following entomologists who supplied them with the flies bearing Labouobeniales and/or identified the host insects: R. Andrade (Portugal), G. Cocks (Australia), J. Deeming (Great Britain), Martin J. Ebejer (Great Britain), Marina Krivosheina (Russia), Iain MacGowan (Great Britain), Thomas Pape (Denmark), Alfio Raspi (Italy), N. Vikhrev (Russia), M. von Tschirnhaus (Germany). H.B. Lee was supported by the Graduate Program for the Undiscovered Taxa of Korea, and in part by the Project on Survey and Discovery of Indigenous Fungal Species of Korea funded by NIBR and Project on Discovery of Fungi from Freshwater and Collection of Fungarium funded by NNIBR of the Ministry of Environment (MOE), and in part carried out with the support of Cooperative Research Program for Agriculture Science and Technology Development (PJ013744), Rural Development Administration, Republic of Korea. This work was in part supported by the BK21 plus program through the National Research Foundation (NRF) funded by the Ministry of Education of Korea. Jian-Kui Liu thanks the National Natural Science Foundation of China (NSFC 31600032). Marcela Cáceres thanks the CNPq (Conselho Nacional de Desenvolvimento Cientı́fico e Tecnológico) for a research grant (309058/2015-5), funding for collecting trips (401186/2014-8), and a collaborative project with RL as Special Visiting Professor (314570/2014-4). Funding for phylogenetic work on Graphidaceae was provided by a grant from the National Science Foundation (NSF) to The Field Museum: DEB-1025861 ‘‘ATM – Assembling a taxonomic monograph: The lichen family Graphidaceae’’; PI Thorsten Lumbsch, CoPI Robert Lücking. Armin Mangold is thanked for providing two photographs of the type material of Thelotrema annulatum. Vanessa P. Abreu, André W. C. Rosado and Olinto L. Pereira thank the CAPES, CNPq, and FAPEMIG for financial support and ICMBio/FLONA-Paraopeba for providing facilities and permits for the exploration surveys of the mycodiversity in their protected areas. We would like to thank the Instituto Chico Mendes de Conservação da Biodiversidade (ICMBio) of the Parque Montanhas do Tumucumaque and the Instituto Brasileiro de Meio Ambiente (IBAMA) for support during the field trips; and Msc. RLM Alvarenga for the figures. We thank the Graduate Program for the Biodiversity and Biotechnology Network of the Legal Amazon (UFPA-MPEG, Brazil); the Museu Paraense Emı́lio Goeldi (MPEG), the Universidade do Estado do Amapa (UEAP) and the Universidade Federal de Pernambuco (UFPE) for the logistical support of its laboratories and herbaria. Further, we acknowledge the Conselho Nacional de Desenvolvimento Cientı́fico Programa de Capacitação for the scholarship to AMSS (Programa de Capacitação Institucional 303073/2018-7); CNPq (Sisbiota 563342/2010-2, PROTAX 562106/2010-3), and FACEPE (APQ 0788-2.03/12) for funding this research. T.G.L. Oliveira, C.M. SouzaMotta, J.D.P. Bezerra, and O.M.C. Magalhães are grateful to the Conselho Nacional de Desenvolvimento Cientı́fico e Tecnológico (CNPq), Coordenação de Aperfeiçoamento de Pessoal de Nı́vel Superior (CAPES) and Fundação de Amparo à Ciência e Tecnologia de Pernambuco (FACEPE) for financial support and/or scholarships; to Dr. Jarcilene S. Almeida Cortez, Dr. Alexandre R. Machado and Greicilene M. R. Albuquerque for their help during the execution of this study, and also to the Instituto Fazenda Tamanduá (Tamanduá Farm) for support during the collections of plant material. PN Singh and Sanjay K Singh thanks to Director Agharkar Research Institute, Pune, India for providing facilities. The technical help in preparing photo-plates by Mr. S.B. Gaikwad is acknowledged. Martina Réblová acknowledges support by a long-term research development project No. RVO 67985939 of the Czech Academy of Sciences, Institute of Botany. José Ewerton F. dos Santos and Renan N. Barbosa acknowledgment scholarship and financial support from Conselho Nacional de Pesquisa e Desenvolvimento Cientifico (CNPq). We would like to acknowledge the URM Herbarium and URM Culture Collection staff and Professor Dr. Roger Fagner R. de Melo for taxonomy assistance. Guo-Jie Li and Rui-Lin Zhao would like to thank National Natural Science Foundation of China 123 222 (Project IDs GJL: 31500013, RLZ: 31470152 and 31360014) for financial support. The material of Plectocarpon on S. tricosa was studied for the Galapagos Lichen Inventory, a joint project of the Charles Darwin Foundation (CDF) and the Galapagos National Park (DPNG), part of a national biodiversity assessment ‘‘Biodiversidad Genética del Ecuador’’ led by the Instituto Nacional de Biodiversidad del Ecuador (INABIO). The authors are indebted to our colleagues and collaborators at these institutions, especially Daniel Lara Solı́s, Galo Quedaza and Victor Carrión (DPNG), Arturo Izurieta and Maria-José Barragan Paladines (CDF), and Diego Inclán and Rosa Batallas (INABIO) for research permits and logistical support. Monika Dayarathne would like to thank Thailand Research Fund (TRF) Grant No. MRG6080089 entitled; Taxonomy and phylogeny of foliar fungi from Mangrove and to Dr. Putarak Chomnunti. Saranyaphat Boonmee would like to thank the Thailand Research Fund (No. TRG6180001), the National Research Council of Thailand (No. 61215320023) and Plant Genetic Conservation Project under the Royal Initiation of Her Royal Highness Princess Maha Chakri Sirindhorn-Mae Fah Luang University. Armin Mešić and Zdenko Tkalčec are grateful to Croatian Science Foundation for their financial support under the project HRZZ-IP-2018-01-1736 (ForFungiDNA) and thank to Milan Čerkez for collecting specimens of coprinoid taxa. Chaynard Phukhamsakda would like to thank the Royal Golden Jubilee PhD Program under Thailand Research Fund (RGJ) for a personal grant to C. Phukhamsakda (The scholarship no. PHD/0020/ 2557 to study towards a PhD). Lei Cai acknowledges China-Thailand Joint Lab on Microbial Biotechnology (Most KY201701011) for financial support. Rungtiwa Phookamsak thanks CAS President’s International Fellowship Initiative (PIFI) for young staff (Grant No. 2019FYC0003), the Research Fund from China Postdoctoral Science Foundation (Grant No. Y71B283261), the Yunnan Provincial Department of Human Resources and Social Security (Grant No. Y836181261), and National Science Foundation of China (NSFC) project code 31850410489 for financial support. Putarak Chomnunti would like to thank the National Research Council of Thailand (Grant No. 256108A3070006) for financial support. Ting-Chi Wen and Yuanpin Xiao would like to thank the National Natural Science Foundation of China (No. 31760014) and the Science and Technology Foundation of Guizhou Province (No. [2016]2863). This research work was partially supported by Chiang Mai University. TB Gibertoni, AMS Soares, HL Plautz-Jr, HMP Sotão and WKS Xavier would like to thank the Instituto Chico Mendes de Conservação da Biodiversidade of the Parque Nacional Montanhas do Tumucumaque and the Instituto Brasileiro de Meio Ambiente for support during the field trips; the Graduate Program for the Biodiversity and Biotechnology Network of the Legal Amazon (UFPA-MPEG), the Museu Paraense Emı́lio Goeldi (MPEG), the Universidade do Estado do Amapa and the Universidade Federal de Pernambuco for the logistical support of their laboratories and herbaria; CNPq for the scholarship of AMSS (Programa de Capacitação Institucional 303073/2018-7); and CNPq (Sisbiota 563342/2010-2, PROTAX 562106/2010-3) and FACEPE (APQ 0788-2.03/12) for funding this research; and Msc. RLM Alvarenga for the figures. B. Buyck thanks the ATM of the Paris’ Museum and ‘‘l’Institut Ecologie et Environnement’’ (CNRS-INEE) for funding the field trip with Shelly Masi to Africa; Shelly is thanked for all the practical help and sharing her experience. Terence Fuh and the staff of the Primate Habituation Programme of the Dzanga-Ndoki National Park of the ‘‘Réserve spéciale de Foret Dense de Dzanga-Sangha’’ at Bayanga, as well as all staff, Aka guides and visitors of the Bai hakou field station for logistical support, field assistance and the very enjoyable company during our stay. This research was made possible through research permit 034/MENESR/ DIRCAB/DGESRSTI/DRSTSPI/SSSTI/16 from the ‘‘Ministère de l’éducation nationale, de l’enseignement supérieur et de la recherche scientifique’’ of the Central African Republic. B. Buyck & V. Hofstetter are grateful to the staff members of CNRE in Antananarivo for logistic support and field assistance; while field work in Madagascar was financed in part by the National Geographic Society (grants 6365-98, 123 Fungal Diversity (2019) 96:1–242 7921-05) and in more recent years by the ATM-project ‘‘Past and present biodiversity’’ of the Muséum national d’histoire naturelle (Dirs. Ph. Janvier and S. Peigné). R Jeewon thanks University of Mauritius for research support. Napalai Chaiwan would like to thank the Thailand Research Fund (PHD60K0147). The material of Plectocarpon on S. tricosa was studied for the Galapagos Lichen Inventory, a joint project of the Charles Darwin Foundation (CDF) and the Galapagos National Park (DPNG), part of a national biodiversity assessment ‘‘Biodiversidad Genética del Ecuador’’ led by the Instituto Nacional de Biodiversidad del Ecuador (INABIO). The authors are indebted to our colleagues and collaborators at these institutions, especially Danny Rueda, Daniel Lara Solı́s, Galo Quedaza and Victor Carrión (DPNG), Arturo Izurieta and Maria-José Barragan Paladines (CDF), and Diego Inclán and Rosa Batallas (INABIO) for research permits and logistical support. This publication is contribution number 2248 of the Charles Darwin Foundation for the Galapagos Islands. Danushka S. Tennakoon would like to thank Lakmali Dissanayake and Binu Samarakoon for their support. D.N. Wanasinghe would like to thank CAS President’s International Fellowship Initiative (PIFI) for funding his postdoctoral research (Number 2019PC0008). Peter E. Mortimer and D.N. Wanasinghe thank the National Science Foundation of China and the Chinese Academy of Sciences for financial support under the following grants: 41761144055, 41771063 and Y4ZK111B01. Yusufjon Gafforov thanks CAS President’s International Fellowship Initiative (Grant No. 2018VBB0021) and German Academic Exchange Service Fellowship (Grant No. 57314018) and Ministry of innovative development of the Republic of Uzbekistan (Projects No. P3-2014-0830174425 and PP20170921183) for funding his research projects. Qiu-Ju Shang would like to thank Dr. Julio Mena-Portales, the senior researcher of the Division of Mycology and president of the Scientific Council, Institute of Ecology and Systematic, Havana, Cuba, for his kind support on manuscript writing. PhD students from Mae Fah Laung and Chiang Mai Universities thank the Mushroom Research Foundation for research financial support and PhD Fellowships. Mingkwan Doilom would like to thank the 5th batch of Postdoctoral Orientation Training Personnel in Yunnan Province (Grant No. Y934283261) and the 64th batch of China Postdoctoral Science Foundation (Grant No. Y913082271). Ningguo Liu, Zuo-Yi Liu, Hong-Yan Su, A. Akulov and Huang Zhang are deeply thanked for their kind support on manuscript writing. Jianchu Xu thanks Key Research Program of Frontier Sciences ‘‘Response of Asian mountain ecosystems to global change’’, CAS (Grant No. QYZDYSSW-SMC014). D.N. Wanasinghe thanks the 64th batch of China Postdoctoral Science Foundation (Grant No. Y913083271). Alan JL Phillips acknowledges the support from UID/MULTI/04046/2019 Research Unit grant from FCT, Portugal to BioISI. References Abarenkov K, Henrik Nilsson R, Larsson KH, Alexander IJ, Eberhardt U, Erland S, Høiland K, Kjøller R, Larsson E, Pennanen T, Sen R (2010) The UNITE database for molecular identification of fungi–recent updates and future perspectives. New Phytol 186:281–285 Abdel-Wahab MA, Jones EBG (2000) Three new marine ascomycetes from driftwood in Australia sand dunes. Mycoscience 41:379–388 Abdollahzadeh J, Javadi A, Goltapeh EM, Zare R, Phillips AJL (2010) Phylogeny and morphology of four new species of Lasiodiplodia from Iran. Persoonia 25:1–10 Abdollahzadeh J, Javadi A, Zare R, Phillips AJL (2014) A phylogenetic study of Dothiorella and Spencermartinsia species associated with woody plants in Iran, New Zealand, Portugal and Spain. Persoonia 32:1–12 Fungal Diversity (2019) 96:1–242 Ahmadpour SA, Mehrabi-Koushki M, Farokhinejad R (2017) Neodidymelliopsis farokhinejadii, a new fungal species from dead branches of trees in Iran. Sydowia 69:171–182 Ahti T, Oksanen J (1990) Epigeic lichen communities of taiga and tundra regions. Vegetatio 86:39–70 Alias SA, Jones EBG, Kuthubutheen AJ (1994) Fasciatispora lignicola sp. nov. (Amphisphaeriaceae, Ascomycotina) from drift mangrove wood. Mycotaxon 52:77–82 Alves A, Crous PW, Correia A, Phillips AJL (2008) Morphological and molecular data reveal cryptic speciation in Lasiodiplodia theobromae. Fungal Divers 28:1–13 Alves A, Linaldeddu B, Deidda A, Scanu B, Phillips A (2014) The complex of Diplodia species associated with Fraxinus and some other woody hosts in Italy and Portugal. Fungal Divers 67:143–156 Amalfi M, Decock C (2013) Fomitiporia castilloi sp. nov. and evidence for multiples clades around F. apiahyna in Meso- and South America, representing potential species. Mycologia 105:873–887 Anagnostakis SL (1987) Chestnut blight: the classical problem of an introduced pathogen. Mycologia 79:23–37 Andersen HL, Ekman S (2004) Phylogeny of the Micareaceae inferred from nrSSU DNA sequences. Lichenologist 36:27–35 Andersen HL, Ekman S (2005) Disintegration of the Micareaceae (lichenized Ascomycota): a molecular phylogeny based on mitochondrial rDNA sequences. Mycol Res 109:21–30 Aptroot A (2002) New and interesting lichens and lichenicolous fungi in Brazil. Fungal Divers 9:15–45 Aptroot A, Cáceres MES (2018a) New lichens from Chapada Diamantina, Bahia, Brazil. Bryologist 121:67–79 Aptroot A, Cáceres MES (2018b) New species and new records of lichens from inselbergs and surrounding Atlantic rain forest in the Chapada Diamantina (Bahia, Brazil). Herzogia 31:359–373 Aptroot A, Diederich P, Sérusiaux E, Sipman HJM (1997) Lichens and lichenicolous fungi from New Guinea. Bibl Lichenol 64:1–220 Ariyawansa HA, Tanaka K, Thambugala KM, Phookamsak R, Tian Q, Camporesi E, Hongsanan S, Monkai J, Wanasinghe DN, Mapook A, Chukeatirote E, Kang JC, Xu JC, McKenzie EHC, Jones EBG, Hyde KD (2014) A molecular phylogenetic reappraisal of the Didymosphaeriaceae (= Montagnulaceae). Fungal Divers 68:69–104 Ariyawansa HA, Hyde KD, Jayasiri SC, Buyck B, Chethana KWT, Dai DQ, Dai YC, Daranagama DA, Jayawardena RS, Lücking R, Ghobad-Nejhad M, Niskanen T, Thambugala KM, Voigt K, Zhao RL, Li GJ, Doilom M, Boonmee S, Yang ZL, Cai Q, Cui YY, Bahkali AH, Chen J, Cui BK, Chen JJ, Dayarathne MC, Dissanayake AJ, Ekanayaka AH, Hashimoto A, Hongsanan S, Jones EBG, Larsson E, Li WJ, Li QR, Liu JK, Luo ZL, Maharachchikumbura SSN, Mapook A, McKenzie EHC, Norphanphoun C, Konta S, Pang KL, Perera RH, Phookamsak R, Phukhamsakda C, Pinruan U, Randrianjohany E, Singtripop C, Tanaka K, Tian CM, Tibpromma S, Abdel-Wahab MA, Wanasinghe DN, Wijayawardene NN, Zhang JF, Zhang H, Abdel-Aziz FA, Wedin M, Westberg M, Ammirati JF, Bulgakov TS, Lima DX, Callaghan TM, Callac P, Chang CH, Coca LF, Dal Forno M, Dollhofer V, Fliegerová K, Greiner K, Griffith GW, Ho HM, Hofstetter V, Jeewon R, Kang JC, Wen TC, Kirk PM, Kytövuori I, Lawrey JD, Xing J, Li H, Liu ZY, Liu XZ, Liimatainen K, Lumbsch HT, Matsumura M, Moncada B, Nuankaew S, Parnmen S, de Azevedo Santiago ALCM, Sommai S, Song Y, de Souza CAF, de Souza-Motta CM, Su HY, Suetrong S, Wang Y, Wei SF, Wen TC, Yuan HS, Zhou LW, Réblová M, Fournier J, Camporesi E, Luangsa-ard JJ, Tasanathai K, Khonsanit A, Thanakitpipattana D, Somrithipol S, Diederich P, Millanes AM, Common RS, Stadler M, Yan JY, Li XH, Lee HW, Nguyen TTT, 223 Lee HB, Battistin E, Marsico O, Vizzini A, Vila J, Ercole E, Eberhardt U, Simonini G, Wen HA, Chen XH (2015) Fungal diversity notes 111–252 taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 75:1–248 Ariyawansa HA, Jaklitsch WM, Voglmayr H (2018) Additions to Taiwan fungal flora 1: neomassariaceae fam. nov. Cryptogam, Mycol 39:1–14 Arzanlou M, Groenewald JZ, Gams W, Braun U, Shin HD, Crous PW (2007) Phylogenetic and morphotaxonomic revision of Ramichloridium and allied genera. Stud Mycol 58:57–93 Asamoah A, Atta-Boateng A, Antwi-Boasiako C, Frimpong-Mensah K, Sarfowaa A (2012) Timbers 2: Intsia bijuga. Prota 7:1–6 Asplund J, Kauserud H, Bokhorst S, Lie MH, Ohlson M, Nybakken L (2018) Fungal communities influence decomposition rates of plant litter from two dominant tree species. Fungal Ecol 32:1–8 Ayoubi N, Soleimani MJ, Zare R (2016) Pilidium concavum, causing tan-brown rot on strawberry in Iran. J Plant Pathol 98:667–669 Bahl J (2006) Molecular evolution of three morphologically similar families in the Xylariomycetidae (Apiosporaceae, Clypeosphaeriaceae, Hyponectriaceae). PhD Thesis, The University of Hong Kong Balgrie B (2003) Taints and off-flavours in food. CRC Press, Boca Raton Baral HO, Haelewaters D, Pärtel K (2015) A new attempt to classify the families of the Helotiales. Poster presented at the Second international workshop on ascomycete systematics, Amsterdam, The Netherlands Barbosa RN, Leong SL, Vinnere-Pettersson O, Chen AJ, Souza-Motta CM, Frisvad JC, Samson RA, Oliveira NT, Houbraken J (2017) Phylogenetic analysis of Monascus and new species from honey, pollen and nests of stingless bees. Stud Mycol 86:29–51 Bargagli-Petrucci G (1915) Studi sulla flora miscroscopica della regione boracifera della Toscana. La vegetazione crittogamica nella regione boracifera. Giorn Bot Ital 22:409–411 Barr DJS (1968) Studies on Rhizophydium and Phyctochytrium (Chytridiales). I. Comparative morphology. Can J Bot 47:991–997 Barr ME (1978) The Diaporthales in North America with emphasis on Gnomonia and its segregates. Mycol Mem 7:1–232 Barr ME (1979) A classification of Loculoascomycetes. Mycologia 71:935–957 Barr ME (1980) On the family Tubeufiaceae (Pleosporales). Mycotaxon 12:137–167 Barr ME (1987) New taxa and combinations in the Loculoascomycetes. Mycotaxon 29:501–505 Barr ME (1992) Notes on the Lophiostomataceae (Pleosporales). Mycotaxon 45:191–221 Beeli M (1928) Contribution à l’étude de la Flore Mycologique du Congo. Fungi Goossensiani VI. Bull Soc R Bot Belg 61(1):78–100 Benjamin RK (1970) Laboulbeniales on semiaquatic Hemiptera II. Autophagomyces, Dioicomyces, and Prolixandromyces gen. nov. Aliso J Syst Evol Bot 7:165–182 Benjamin RK (1999) Laboulbeniales on semiaquatic Heteroptera VIII. Monandromyces, a genus based on Autophagomyces microveliae (Laboulbeniales). Aliso J Syst Evol Bot 18:71–91 Berbee ML (1996) Loculoascomycete origins and evolution of filamentous ascomycete morphology based on 18S rRNA gene sequence data. Mol Biol Evol 13:462–470 Berkeley MJ, Curtis MA (1869) Fungi Cubenses (Hymenomycetes). J Linn Soc 10:280–392 Bidartondo MI (2005) The evolutionary ecology of myco-heterotrophy. New Phytol 167:335–352 Bills GF, Platas G, Gams W (2004) Conspecificity of the cerulenin and helvolic acid producing ‘Cephalosporium caerulens’, and the 123 224 hypocrealean fungus Sarocladium oryzae. Mycol Res 108:1291–1300 Binder M, Justo A, Riley R, Salamov A, Lopez-Giraldez F, Sjökvist E, Hibbett DS (2013) Phylogenetic and phylogenomic overview of the Polyporales. Mycologia 105:1350–1373 Boerema GH, Loerakker WM, Hamers ME (1996) Contributions towards a monograph of Phoma (Coelomycetes)—III. 2. Misapplications of the type species name and the generic synonyms of section Plenodomus (Excluded species). Persoonia 16:141–189 Boonmee S, Zhang Y, Chomnunti P, Chukeatirote E, Tsui CKM, Bahkali AH, Hyde KD (2011) Revision of lignicolous Tubeufiaceae based on morphological reexamination and phylogenetic analysis. Fungal Divers 51:63–102 Boonmee S, Rossman AY, Liu JK, Li WJ, Dai DQ, Bhat DJ, Jones EBG, McKenzie EHC, Xu JC, Hyde KD (2014) Tubeufiales, ord. nov., integrating sexual and asexual generic names. Fungal Divers 68:239–298 Boonmee S, D’souza MJ, Luo ZL, Pinruan U, Tanaka K, Su HY, Bhat DJ, Mckenzie E, Jones EBG, Taylor JE, Phillips AJL, Hirayama K, Eungwanichayapant PD, Hyde KD (2016) Dictyosporiaceae fam. nov. Fungal Divers 80:457–482 Boonyuen N, Pang KL (2011) Savoryellales (Hypocreomycetidae, Sordariomycetes): a novel lineage of aquatic ascomycetes inferred from multiple-gene phylogenies of the genera Ascotaiwania, Ascothailandia, and Savoryella. Mycologia 103:1351–1371 Brahmanage RS, Lu YZ, Bhat DJ, Wanasinghe DN, Yan JY, Hyde KD, Boonmee S (2017) Phylogenetic investigations on freshwater fungi in Tubeufiaceae (Tubeufiales) reveals the new genus Dictyospora and new species Chlamydotubeufia aquatica and Helicosporium flavum. Mycosphere 8:917–933 Brummelen J, Kristiansen R (1999) A new species of Boubovia (Pezizales) from the Hvaler Archipelago in Norway. Persoonia 17:265–271 Buyck B, Hofstetter V (2018) Cantharellus subg. Pseudocantharellus (Hydnaceae, Cantharellales) revisited: one epityfication, two new synonyms and one new species. Mycosphere 9:141–148 Buyck B, Mitchell D, Parrent J (2006) Russula parvovirescens sp. nov., a common but ignored species in the eastern United States. Mycologia 98:612–615 Buyck B, Hofstetter V, Eberhardt U, Verbeken A, Kauff F (2008) Walking the thin line between Russula and Lactarius: the dilemma of Russula sect. Ochricompactae. Fungal Divers 28:15–40 Buyck B, Hofstetter V, Verbeken A, Walleyn R (2010) Proposal to conserve Lactarius nom. cons. (Basidiomycota) with conserved type. Taxon 59:447–453 Buyck B, Randrianjohany E, Eyssartier G (2012) Observations on some enigmatic Cantharellus (Cantharellales, Basidiomycota) with lilac-violaceous tints from Africa and Madagascar. Cryptogam, Mycol 33:167–179 Buyck B, Kauff F, Randrianjohany E, Hofstetter V (2015) Sequence data reveal a high diversity of Cantharellus associated with endemic vegetation in Madagascar. Fungal Divers 70:189–208 Buyck B, Henkel TW, Dentinger BTM, Séné O, Hofstetter V (2016a) Multigene sequencing provides a suitable epitype barcode sequences and a precise systematic position for the enigmatic African Cantharellus miniatescens. Cryptogam, Mycol 37:269–282 Buyck B, Olariaga I, Hofstetter V (2016b) Setting the record straight on North American Cantharellus. Cryptogam, Mycol 37:405–417 Buyck B, Olariaga I, Justice J, Lewis D, Hofstetter V (2016c) The dilemma of species recognition in the field when sequence data 123 Fungal Diversity (2019) 96:1–242 are not in phase with phenotypic variability. Cryptogam, Mycol 37:367–390 Buyck B, Duhem B, Das K, Jayawardena RS, Niveiro N, Pereira OL, Prasher IB, Adhikari S, Albertó EO, Bulgakov TS, CastañedaRuiz RF, Hembrom ME, Hyde KD, Lewis DP, Michlig A, Nuytinck J, Parihar A, Popoff OF, Ramirez NA, da Silva M, Verma RK, Hofstetter V (2017) Fungal biodiversity profiles 21–30. Cryptogam Mycol 38:101–146 Buyck B, Henkel TW, Hofstetter V (2018) The Cantharellus isabellinus complex Cantharellales, Hydnaceae) in tropical Africa. Mycosphere 9:1209–1221 Cáceres MES, Mota N Jr, Santos LA, Pereira TA, Aptroot A (2017) New records to Brazil and Southern Hemisphere of corticolous and saxicolous lichens from the semiarid region in Ceará State. Iheringia 72:239–245 Cai L, Zhang K, McKenzie EH, Ho WH, Hyde KD (2002) Acrodictys liputii sp. nov. and Digitodesmium bambusicola sp. nov. from bamboo submerged in Liput River in the Philippines. Nova Hedw 75:525–532 Cai L, Zhang KQ, McKenzie EHC, Hyde KD (2003) Freshwater fungi from bamboo and wood submerged in the Liput River in the Philippines. Fungal Divers 13:1–12 Cai L, Hyde KD, Tsui CKM (2006) Genera of freshwater fungi. Fungal Divers Res Ser 18:1–261 Campbell J, Ferrer A, Raja HA, Sivichai S, Shearer CA (2007) Phylogenetic relationships among taxa in the Jahnulales inferred from 18S and 28S nuclear ribosomal DNA sequences. Can J Bot 85:873–882 Cannon PF, Kirk PM (2007) Fungal families of the world. CABI, Wallingford Cardin L, Vincenot L, Balesdent MH (2009) First Report of Pilidium concavum on Bergenia crassifolia in France. Plant Dis 93:548 Carlucci A, Cibelli F, Lops F, Raimondo ML (2015) Characterization of Botryosphaeriaceae species as causal agents of trunk diseases on grapevines. Plant Dis 99(12):1678–1688 Castaldo D, Rossi W, Sabatini F (2004) Contribution to the knowledge of the Laboulbeniales from Greece. Plant Biosyst 138:261–269 Castañeda Ruı́z RF, Velasquez S, Cano J, Saikawa M, Guarro J (2002) Phaeoisaria aguilerae anam. sp. nov. from submerged wood in Cuba with notes and reflections in the genus Phaeoisaria. Cryptogam, Mycol 23:9–18 Chang HS (1995) Notes on Taiwan dematiaceous hyphomycetes, some species of the genera Exserticlava, Craspedodidymum and Hermatomyces. Bot Bull Acad Sin (Taiwan) 36:243–246 Checa J (1997) Annotated list of the Lophiostomataceae and Mytilinidiaceae (Dothideales, Ascomycotina) reported from the Iberian Peninsula and Balearic Islands. Mycotaxon 63:467–491 Chen H, Cui BK (2017) Multi-locus phylogeny and morphology reveal five new species of Fomitiporia (Hymenochaetaceae) from China. Mycol Prog 16:687–701 Chen SF, Morgan DP, Michailides TJ (2013) First report of Phoma fungicol aassociated with stem canker and fruit blight of pistachio in Arizona. J Plant Pathol 95:451 Chen Q, Jiang JR, Zhang GZ, Cai L, Crous PW (2015) Resolving the Phoma enigma. Stud in Mycol 82:137–217 Chen Q, Hou LW, Duan WJ, Crous PW, Cai L (2017) Didymellaceae revisited. Stud Mycol 87:105–159 Cheng XL, Li W, Zhang TY (2014) A new species of Phaeoisaria from intertidal marine sediment collected in Weihai, China. Mycotaxon 127:17–24 Chesters CGC, Bell A (1970) Studies in the Lophiostomataceae Sacc. Myc Pap 120:1–55 Chevallier FF (1826) Flore Générale des Environs de Paris, vol 1. Ferra Jeune, Paris, pp 1–674 Fungal Diversity (2019) 96:1–242 Clements FE, Shear CL (1931) Genera of fungi 2, vols i–vii. H.W. Wilson, New York Conceição AA, Pirani JR, Meirelles ST (2007) Floristics, structure and soil of insular vegetation in four quartzite-sandstone outcrops of ‘‘Chapada Diamantina’’, Northeast Brazil. Braz J Bot 30:641–656 Costa PMO, Barbosa MA, Silva GVRD, Sosa D, Pérez-Martinez S, Castañeda-Ruiz RF, Malosso E (2018) Castanediella diversispora sp. nov. from the Brazilian Atlantic Forest. Mycotaxon 133:63–69 Crane JL, Miller AN (2016) Studies in genera similar to Torula: Bahusaganda, Bahusandhika, Pseudotorula, and Simmonsiella gen. nov. IMA Fungus 7:29–45 Crespo A, Kauff F, Divakar PK, del Prado R, Pérez-Ortega S, Amo de Paz G, Ferencova Z, Blanco O, Roca-Valiente B, Núñez-Zapata J, Cubas P, Argüello A, Elix JA, Esslinger TL, Hawksworth DL, Millanes A, Molina MC, Wedin M, Ahti T, Aptroot A, Barreno E, Bungartz F, Calvelo S, Candan M, Cole M, Ertz D, Goffinet B, Lindblom L, Lücking R, Lutzoni F, Mattsson J-E, Messuti MI, Miadlikowska J, Piercey-Normore M, Rico VJ, Sipman HJM, Schmitt I, Spribille T, Thell A, Thor G, Upreti DK, Lumbsch HT (2010) Phylogenetic generic classification of parmelioid lichens (Parmeliaceae, Ascomycota) based on molecular, morphological and chemical evidence. Taxon 59:1735–1753 Crous PW, Braun U, Groenewald JZ (2007) Mycosphaerella is polyphyletic. Stud Mycol 58:1–32 Crous PW, Summerell BA, Carnegie AJ, Wingfield MJ, Groenewald JZ (2009) Novel species of Mycosphaerellaceae and Teratosphaeriaceae. Persoonia 23:119–146 Crous PW, Verkley GJM, Christensen M, Castañeda-Ruiz RF, Groenewald JZ (2012) How important are conidial appendages? Persoonia 28:126–137 Crous PW, Shivas RG, Quaedvlieg W, van der Bank M, Zhang Y, Summerell BA, Guarro J, Wingfield MJ, Wood AR, Alfenas AC, Braun U, Cano-Lira JF, Garcı́a D, Marin-Felix Y, Alvarado P, Andrade JP, Armengol J, Assefa A, den Breeÿen A, Camele I, Cheewangkoon R, De Souza JT, Duong TA, Esteve-Raventós F, Fournier J, Frisullo S, Garcı́a-Jiménez J, Gardiennet A, Gené J, Hernández-Restrepo M, Hirooka Y, Hospenthal DR, King A, Lechat C, Lombard L, Mang SM, Marbach PAS, Marincowitz S, Marin-Felix Y, Montaño-Mata NJ, Moreno G, Perez CA, Pérez Sierra AM, Robertson JL, Roux J, Rubio E, Schumacher RK, Stchigel AM, Sutton DA, Tan YP, Thompson EH, van der Linde E, Walker AK, Walker DM, Wickes BL, Wong PTW, Groenewald JZ (2014) Fungal planet description sheets: 214–280. Persoonia 32:184–306 Cruywagen EM, Slippers B, Roux J, Wingfield MJ (2016) Phylogenetic species recognition and hybridization in Lasiodiplodia: a case study on species from baobabs. Fungal Biol 121:420–436 Cui BK, Du P, Dai YC (2011) Three new species of Inonotus (Basidiomycota, Hymenochaetaceae) from China. Mycol Prog 10:107–114 Czarnota P, Guzow-Krzemińska B (2010) A phylogenetic study of the Micarea prasina group shows that Micarea micrococca includes three distinct lineages. Lichenologist 42:7–21 Dai YC (2010) Hymenochaetaceae (Basidiomycota) in China. Fungal Divers 45:131–343 Dai YC (2012) Polypore diversity in China with an annotated checklist of Chinese polypores. Mycoscience 53:49–80 Dai YC, Cui BK, Yuan HS, Li BD (2007) Pathogenic wood-decaying fungi in China. For Pathol 37:105–120 Dai YC, Yang ZL, Cui BK, Yu CJ, Zhou LW (2009) Species diversity and utilization of medicinal mushrooms and fungi in China (Review). Int J Med Mushrooms 11:287–302 225 Dai DQ, Tang LZ, Liu C, Wang HB, Hyde KD (2018) Studies on Parmulariaceae I. A phylogeny based on available sequence data; introducing Parmulariales ord. nov., and Hemigraphaceae. Melaspileellaceae and Stictographaceae fam. nov. Phytotaxa 369:63–79 Daranagama DA, Hyde KD, Sir EB, Thambugala KM, Tian Q, Samarakoon MC, McKenzie EH, Jayasiri SC, Tibpromma S, Bhat JD, Liu X (2018) Towards a natural classification and backbone tree for Graphostromataceae, Hypoxylaceae, Lopadostomataceae and Xylariaceae. Fungal Divers 88:1–165 Das K, Atri NS, Buyck B (2013) Three new species of Russula (Russulales) from India. Mycosphere 4:722–732 Das K, Ghosh A, Bhatt RP, Chakraborty D, Hofstetter V, Buyck B (2017a) Fungal biodiversity profiles 41–50. Cryptogam Mycol 38:527–547 Das K, Ghosh A, Chakraborty D, Li J, Qiu L, Baghela A, Halama M, Hembrom ME, Mehmood T, Parihar A, Pencakowski B, Bielecka M, Reczyńska K, Sasiela D, Singh U, Song Y, Świerkosz K, Szcze˛śniak K, Uniyal P, Zhang J, Buyck B (2017b) Fungal biodiversity profiles 31–40. Cryptogam Mycol 38:353–406 Das K, Rossi W, Leonardi M, Ghosh A, Bera I, Hembrom ME, Bajpai R, Nayaka JS, Upreti DK, Wang X-H, Hofstetter V, Buyck B (2018a) Fungal biodiversity profiles 61–70. Cryptogam, Mycol 39:381–418 Das K, Ghosh A, Baghela A, Avchar R, Bhatt RP, Buyck B (2018b) Morphological and phylogenetic evidence for Russula butyroindica sp. nov. (Russulaceae) from the Indian Himalaya. Nova Hedw 106:295–303 Dayarathne MC, Maharachchikumbura SSN, Jones EBG, Wei D, Devadatha B, Yang J, Ekanayake H, De Silva W, Sarma VV, AlSadi AM, Khongphinitbunjong K, Hyde KD, Zhao RL (2019a) Phylogenetic revision of Savoryellaceae and evidence for its ranking as a subclass. Front Microbiol 10:840 Dayarathne MC, Wanasinghe DN, Devadatha B, Abeywickrama P, Jones EBG, Chomnunti P, Sarma VV, Hyde KD, Lumyong S, McKenzie EHC (2019b) Modern taxonomic approaches to identifying diatrypaceous fungi from marine habitats, a novel genus Halocryptovalsa and expansion of Halodiatrype. Cryptogam Mycol (submitted) de Gruyter JD, Woudenberg JHC, Aveskamp MM, Verkley GJM, Groenewald JZ, Crous PW (2010) Systematic reappraisal of species in Phoma section Paraphoma, Pyrenochaeta and Pleurophoma. Mycologia 102:1066–1081 de Gruyter JD, Woudenberg JHC, Aveskamp MM, Verkley GJM, Groenewald JZ, Crous PW (2013) Redisposition of Phoma-like anamorphs in Pleosporales. Stud in Mycol 75:1–36 De Almeida DAC, Gusmão LFP, Miller AN (2016) Taxonomy and molecular phylogeny of Diatrypaceae (Ascomycota, Xylariales) species from the Brazilian semi-arid region, including four new species. Mycol Prog 15:1–27 De Hoog GS (2014) Ecology and phylogeny of black yeast-like fungi: diversity in unexplored habitats. Fungal Divers 65:1–2 De Hoog GS, Guarro J, Gene J, Figueras MJ (2000) Atlas of clinical fungi, vol 1. Centraalbureau voor Schimmelcultures, Utrecht, p 432p De Kesel A, Amalfi M, Ngoy BKW, Yorou NS, Raspé O, Degreef J, Buyck B (2016) New and interesting Cantharellus from tropical Africa. Cryptogam, Mycol 37:283–327 De Notaris G (1846) Frammenti lichenografici di un lavoro inedito. Giornale Bot Ital 2(1):174–224 De Silva N, Lumyong S, Hyde KD, Bulgakov T, Phillips AJL, Yan JY (2016) Mycosphere essays 9: defining biotrophs and hemibiotrophs. Mycosphere 7:545–559 Dearnaley JDW (2007) Further advances in orchid mycorrhizal research. Mycorrhiza 17:475–486 123 226 Deighton FC (1974) Four synnematous hyphomycetes. Trans Br Mycol Soc 62:243–252 Delgado G, Koukol O, Cáceres O, Piepenbring M (2017) The phylogenetic placement of Ernakulamia cochinensis within Pleosporales (Dothideomycetes, Ascomycota). Cryptogam, Mycol 38:435–452 Descals E, Nawawi A, Webster J (1976) Developmental studies in Actinospora and three similar aquatic hyphomycetes. Trans Br Mycol Soc 67:207–222 Devadatha B, Sarma V, Wanasinghe D, Hyde KD, Jones EBG (2017) Introducing the new Indian mangrove species, Vaginatispora microarmatispora (Lophiostomataceae) based on morphology and multigene phylogenetic analysis. Phytotaxa 329:139–149 Diederich P, Lawrey JD, Ertz D (2018) The 2018 classification and checklist of lichenicolous fungi, with 2000 non-lichenized, obligately lichenicolous taxa. Bryologist 121:340–425 Dissanayake AJ, Phillips AJL, Hyde KD, Li XH (2016) Botryosphaeriaceae: current status of genera and species. Mycosphere 7:1001–1073 Dissanayake AJ, Camporesi E, Hyde KD, Yan JY, Li XH (2017a) Saprobic Botryosphaeriaceae, including Dothiorella italica sp. nov., associated with urban and forest trees in Italy. Mycosphere 8:1157–1176 Dissanayake AJ, Phillips AJL, Hyde KD, Yan JY, Li XH (2017b) The current status of species in Diaporthe. Mycosphere 8:1106–1156 Doidge EM (1942) A revision of South African Microthyriaceae. Bothalia 4(2):273–420 Doilom M, Liu JK, Jaklitsch WM, Ariyawansa H, Wijayawardene NN, Chukeatirote E, Zhang M, McKenzie EH, Geml J, Voglmayr H, Hyde KD (2013) An outline of the family Cucurbitariaceae. Sydowia 65:167–192 Doilom M, Dissanayake AJ, Wanasinghe DN, Boonmee S, Liu JK, Bhat DJ, Taylor JE, Bahkali AH, McKenzie EHC, Hyde KD (2017) Microfungi on Tectona grandis (teak) in Northern Thailand. Fungal Divers 82:107–182 Doilom M, Hyde KD, Phookamsak R, Dai DQ, Tang LZ, Hongsanan S, Chomnunti P, Boonmee S, Dayarathne MC, Li WJ, Thambugala KM, Perera RH, Daranagama DA, Norphanphoun C, Konta S, Dong W, Ertz D, Phillips AJL, McKenzie EHC, Vinit K, Ariyawansa HA, Jones EBG, Mortimer PE, Xu JC, Promputtha I (2018) Mycosphere Notes 225–274: types and other specimens of some genera of Ascomycota. Mycosphere 9:647–754 Dong W, Hyde KD, Bhat DJ, Zhang H (2018) Introducing Aculeata aquatica gen. et sp. nov., Minimelanolocus thailandensis sp. nov. and Thysanorea aquatica sp. nov. (Herpotrichiellaceae, Chaetothyriales) from freshwater in northern Thailand. Mycol Prog 17:617–629 Donk MA (1948) Notes on Malesian fungi. I. Bull Bot Gard Buitenzorg 17:473–482 Donk MA (1971) Progress in the study of the classification of the higher Basidiomycetes. In: Petersen RH (ed) Evolution in the higher basidiomycetes. University of Tennessee Press, Knoxville, pp 3–25 Doweld A (2001) Prosyllabus Tracheophytorum. Geos, Moscow Dubey R, Neelima AM (2013) Kamalomyces mahabaleshwarensis sp. nov. (Tubeufiaceae) from the Western Ghats, India. Mycosphere 4:760–763 Dutta AK, Paloi S, Pradhan P, Acharya K (2015) A new species of Russula (Russulaceae) from India based on morphological and molecular (ITS and LSU sequence) data. Turk J Bot 39:850–856 Ebrahimi L, Fotouhifar K (2016) Identification of some fungi accompanying the scab symptoms in Iran. Mycol Iran 3:25–37 Ekman S, Svensson M (2014) Brianaria (Psoraceae), a new genus to accommodate the Micarea sylvicola group. The Lichenologist 46(3):285–294 123 Fungal Diversity (2019) 96:1–242 Ekanayaka AH, Ariyawansa HA, Hyde KD, Jones EBG, Daranagama DA, Phillips AJL, Hongsanan S, Jayasiri SC, Zhao Q (2017) Discomycetes: the apothecial representatives of the phylum Ascomycota. Fungal Divers 87:237–298 Ekanayaka AH, Hyde KD, Jones EBG, Zhao Q (2018) Taxonomy and phylogeny of operculate discomycetes: Pezizomycetes. Fungal Divers 90:161–243 Ellis MB (1971) Dematiaceous hyphomycetes. Commonwealth Mycological Institute, Kew Ellis MB (1972) Dematiaceous Hyphomycetes. XI. Mycol Pap 131:1–25 Ellis MB (1976) More dematiaceous hyphomycetes. Commonwealth Mycological Institute, Kew Eriksson OE (2006) Outline of Ascomycota—2006. Myconet 12:1–82 Eriksson OE, Hawksworth DL (1986) An alphabetical list of the generic names of ascomycetes. CAB International Mycological Institute, Wallingford Eriksson OE, Hawksworth DL (1987) Notes on ascomycete systematics. Nos. 464–551 Eriksson OE, Winka K (1997) Supraordinal taxa of Ascomycota. Myconet 1:1–16 Ertz D, Tehler A (2011) The phylogeny of Arthoniales (Pezizomycotina) inferred from nucLSU and RPB2 sequences. Fungal Divers 49:47–71 Ertz D, van den Boom P (2012) Plectocarpon dirinariae (Arthoniales), a new lichenicolous species from Cape Verde. Lichenologist 44:591–593 Ertz D, Christnach C, Wedin M, Diederich P (2005) A world monograph of the genus Plectocarpon. Biblioth Lichenol 91:1–155 Etayo J (2017) Hongos liquenı́colas de Ecuador. Opera Lilloana 50:1–535 Eyssartier G (2001) Vers une monographie du genre Cantharellus Adans.:Fr. Dissertation, National History Museum, Paris Farr DF, Rossman AY (2019) Fungal databases. U.S. National Fungus Collections, ARS, USDA. https://nt.ars-grin.gov/fungaldata bases/. Accessed 8 Mar 2019 Ferrer A, Sivichai S, Shearer CA (2007) Megalohypha, a new genus in the Jahnulales from aquatic habitats in the tropics. Mycologia 99:456–460 Fiasson JL, Niemelä T (1984) The Hymenochaetales: a revision of the European poroid taxa. Karstenia 24:14–28 Finlay RD (2008) Ecological aspects of mycorrhizal symbiosis: with special emphasis on the functional diversity of interactions involving the extraradical mycelium. J Exp Bot 59:1115–1126 Fournier J, Raja HA, Shearer CA (2015) Freshwater Ascomycetes: Jahnula purpurea (Jahnulales, Dothideomycetes), a new species on submerged wood from Martinique Island, Lesser Antilles. MycoKeys 9:29–36 Frey W, Hurka H, Oberwinkler F (1977) Beiträge zur Biologie der niederen Pflanzen. Gustav Fischer Verlag, Stuttgart, pp 1–233 Fries E (1838) Epicrisis Systematis Mycologici. Typographia Academica, Uppsala Frisch A, Kalb K, Grube M (2006) Contributions towards a new systematics of the lichen family Thelotremataceae. Biblioth Lichenol 92:1–556 Frisch A, Thor G, Ertz D, Grube M (2014) The Arthonialean challenge: restructuring Arthoniaceae. Taxon 63:727–744 Fröhlich J, Hyde KD (2000) Palm Microfungi. Fungal Divers Res. Ser. 3:1-393 Frøslev TG, Jeppesen TS, Læssøe T (2006) Seven new calochroid and fulvoid species of Cortinarius. Mycol Res 110:1046–1058 Frøslev TG, Brandrud TE, Jeppesen TS (2006) New species and combinations in Cortinarius subgenus Phlegmacium section Calochroi. Mycotaxon 97:367–377 Fungal Diversity (2019) 96:1–242 Funch RR, Harley RM, Funch LS (2009) Mapping and evaluation of the state of conservation of the vegetation in and surrounding the Chapada Diamantina National Park, NE Brazil. Biota Neotrop 9:21–30 Galloway D (2006) [Review of:] Ertz D, Christnach C, Wedin M, Diederich P: A world monograph of the genus Plectocarpon (Roccellaceae, Arthoniales). Bibliotheca Lichenologica 91. J. Cramer, Berlin and Stuttgart. 155 pages. 2005. Australas Lichenol 58:40–42 Gams W, Hawksworth DL (1975) The identity of Acrocylindrium oryzae Sawada and a similar fungus causing sheath-rot of rice. Kavaka 3:57–61 Geiser DM, Gueidan C, Miadlikowska J, Lutzoni F, Kauff F, Hofstetter V, Fraker E, Schoch CL, Tibell L, Untereiner WA, Aptroot A (2006) Eurotiomycetes: Eurotiomycetidae and Chaetothyriomycetidae. Mycologia 98:1053–1064 Geng WL, Hu P, Ma Z, Zhao XY, Wei YM (2012) First report of Pilidium concavum causing tan-brown leaf spot on strawberry in China. Plant Dis 96:1377 Geng K, Zhang B, Song Y, Hyde KD, Kang JC, Wang Y (2013) A new species of Pestalotiopsis from leaf spots of Licuala grandis from Hainan, China. Phytotaxa 88:49–54 Giraldo A, Gené J, Sutton DA, Madrid H, De Hoog GS, Cano J, Decock C, Crous PW, Guarro J (2015) Phylogeny of Sarocladium (Hypocreales). Persoonia 34:10–24 Glawe DA, Rogers JD (1984) Diatrypaceae in the Pacific Northwest. Mycotaxon 20:401–446 Goh TK, Ho WH, Hyde KD, Umali TE (1997) New records and species of Sporoschisma and Sporoschismopsis from submerged wood in the tropics. Mycol Res 101:1295–1307 Goh TK, Hyde KD, Tsui KM (1998a) The hyphomycete genus Acrogenospora, with two new species and two new combinations. Mycol Res 102:1309–1315 Goh TK, Ho WH, Hyde KD, Whitton SR, Umali TE (1998b) New records and species of Canalisporium (Hyphomycetes), with a revision of the genus. Can J Bot 76:142–152 Goldmann L, Weir A (2018) Molecular phylogeny of the Laboulbeniomycetes (Ascomycota). Fungal Biol 122:87–100 Goonasekara ID, Jayawardene RS, Saichana N, Hyde KD (2018) Checklist of microfungi on grasses in Thailand (excluding bambusicolous fungi). Asian J Mycol 1:88–105 Goos RD, Brooks RD, Lamore BJ (1977) An undescribed hyphomycete from wood submerged in a Rhode Island stream. Mycologia 69:280–286 Guatimosim E, Firmino A, Bezerra J, Pereira O, Barreto R, Crous P (2015) Towards a phylogenetic reappraisal of Parmulariaceae and Asterinaceae (Dothideomycetes). Persoonia 35:230–241 Gueidan C, Aptroot A, da Silva Cáceres ME, Badali H, Stenroos S (2014) A reappraisal of orders and families within the subclass Chaetothyriomycetidae (Eurotiomycetes, Ascomycota). Mycol Prog 13:1027–1039 Guindon S, Gascuel O (2003) A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol 52:696–704 Guindon S, Dufayard JF, Lefort V, Anisimova M, Hordijk W, Gascuel O (2010) New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol 59:307–321 Güngör H, Güngör MY, Solak MH (2014) Ascodesmidaceae J. Schröt. a new family (Ascomycota) record for the Turkish mycota. Biodivers Conserv 7:115–116 Guo JY, Karunarathna SC, Mortimer PE, Xu JC, Hyde KD (2014) Phylogenetic diversity of Russula from Xiaozhongdian, Yunnan, China, inferred from internal transcribed spacer sequence data. Chiang Mai J Sci 41:811–821 227 Haase G, Sonntag L, Van de Peer Y, Uijthof JMJ, Podbielski A, Melzer-Krick B (1995) Phylogenetic analysis of ten black yeast species using nuclear small subunit rRNA gene sequences. Anton Leeuw 68:19–33 Haelewaters D, Pfliegler WP, Gorczak M, Pfister DH (2019) Birth of an order: comprehensive molecular phylogenetic study excludes Herpomyces (Fungi, Laboulbeniomycetes) from Laboulbeniales. Mol Phylogenet Evol 133:286–301 Hafellner J, Kalb K (1995) Studies in Trichotheliales ordo novus. Biblioth Lichenol 57:161–186 Hale ME Jr (1980) Generic delimitation in the lichen family Thelotremataceae. Mycotaxon 11:130–138 Hansen K, Pfister DH (2006) Systematics of the Pezizomycetes—the operculate discomycetes. Mycologia 98:1029–1040 Hansen K, Perry BA, Dranginis AW, Pfister DH (2013) A phylogeny of the highly diverse cup-fungus family Pyronemataceae (Pezizomycetes, Ascomycota) clarifies relationships and evolution of selected life history traits. Mol Phylogenet Evol 67:311–335 Hansford GC (1946) The foliicolous ascomycetes, their parasites and associated fungi. Mycol Pap 15:1–240 Hashimoto A, Matsumura M, Hirayama K, Tanaka K (2017) Revision of Lophiotremataceae (Pleosporales Dothideomycetes): Aquasubmersaceae, Cryptocoryneaceae and Hermatomycetaceae fam. nov. Persoonia 39:51–73 Hashimoto A, Hirayama K, Takahashi H, Matsumura M, Okada G, Chen CY, Huang JW, Kakishima M, Ono T, Tanaka K (2018) Resolving the Lophiostoma bipolare complex: generic delimitations within Lophiostomataceae. Stud Mycol 90:161–189 Hawksworth DL (1984) Observations on Jahnula Kirschst., a remarkable aquatic pyrenomycete. Sydowia 37:43–46 Hawksworth DL (2001) The magnitude of fungal diversity: the 1 5 million species estimate revisited. Mycol Res 105:1422–1432 Hawksworth DL, Eriksson OE (1987) Notes on ascomycete systematics. Nos 225–463. Syst Ascomycetum 6:111–165 Hawksworth DL, Galloway DJ (1984) The identity of Plectocarpon Fée, and its implications for Lichenomyces pseudocyphellaria and the typification of Sticta delisea. Lichenologist 16:85–89 Hawksworth DL, Lücking R (2017) Fungal diversity revisited: 2.2 to 3.8 million species. Microbiol Spectr 5:FUNK-0052-2016 He SH, Dai YC (2012) Taxonomy and phylogeny of Hymenochaete and allied genera of Hymenochaetaceae (Basidiomycota) in China. Fungal Divers 56:77–93 Heinemann P (1958) Champignons récoltés au Congo Belge par Madame Gossens-Fontana. III. Cantharellineae. Bull Jard Bot Brux 28:385–438 Heinemann P (1959) Cantharellineae. Flore Iconogr Champignons Congo 8:153–165 Herrera C, Rossman A, Samuels G, Lechat C, Chaverri P (2013) Revision of the genus Corallomycetella with Corallonectria gen. nov. for C. jatrophae (Nectriaceae, Hypocreales). Mycosystema 32:518–544 Hernández-Restrepo M, Gené J, Castañeda-Ruiz RF, Mena-Portales J, Crous PW, Guarro J (2017) Phylogeny of saprobic microfungi from Southern Europe. Stud Mycol 86:53–97 Hibbett DS, Binder M (2002) Evolution of complex fruiting-body morphologies in homobasidiomycetes. Proc R Soc Lond B 269:1963–1969 Hibbett DS, Binder M, Bischoff JF, Blackwell M, Cannon PF, Eriksson OE, Huhndorf S, James T, Kirk PM, Lücking R, Lumbsch HT (2007) A higher-level phylogenetic classification of the fungi. Mycol Res 11:509–547 Hibbett DS, Bauer R, Binder M, Giachini AJ, Hosaka K, Justo A, Larsson E, Larsson KH, Lawrey JD, Miettinen O, Nagy LG (2014) 14 Agaricomycetes. Systematics and evolution. Springer, Berlin, pp 373–429 123 228 Hidayat I, Meeboon J, To-Anun C (2007) Anthostomella and Fasciatispora species (Xylariaceae) from palms including F. ujungkulonensis sp. nov. Mycotaxon 102:347–354 Hirayama K, Tanaka K (2011) Taxonomic revision of Lophiostoma and Lophiotrema based on reevaluation of morphological characters and molecular analyses. Mycoscience 52:401–412 Ho WH, Hyde KD, Hodgkiss IJ (1999) Digitodesmium recurvum, a new species of chirosporous hyphomycete from Hong Kong. Mycologia 91:900–904 Ho WH, Yanna Hyde KD, Hodgkiss IJ (2002) Seasonality and sequential occurrence of fungi onwood submerged in Tai Po Kau forest stream, Hong Kong. Fungal Divers 10:21–43 Holm L (1957) Études taxonomiques sur les Pléosporacées, vol 14. Doctoral dissertation, Acta Universitatis Upsaliensis, pp 1–188 Holm L (1961) Taxonomical notes on Ascomycetes. IV. Notes of Nodulosphaeria Rbh. Sven Bot Tidskr 55:63–80 Holm L, Holm K (1988) Studies in the Lophiostomataceae with emphasis on the Swedish species. Acta Univ Ups Symb Bot Ups 28:1–50 Hongsanan S, Li YM, Liu JK, Hofmann T, Piepenbring M, Bhat JD, Boonmee S, Doilom M, Singtripop C, Tian Q, Mapook A, Zeng XY, Bahkali AH, Xu JC, Mortimer PE, Wu HX, Yang JB, Hyde KD (2014) Revision of genera in Asterinales. Fungal Divers 68:1–68 Hongsanan S, Maharachchikumbura SSN, Hyde KD, Samarakoon MC, Jeewon R, Zhao Q, Al-Sadi AM, Bahkali AH (2017) An updated phylogeny of Sordariomycetes based on phylogenetic and molecular clock evidence. Fungal Divers 84:25–41 Hosagoudar VB, Abraham TK, Biju CK (2001) Re-evaluation of the family Asterinaceae. J Mycopathol Res 39:61–63 Houbraken J, Samson RA (2011) Phylogeny of Penicillium and the segregation of Trichocomaceae into three families. Stud Mycol 70:1–51 Houbraken J, de Vries RP, Samson RA (2014) Modern taxonomy of biotechnologically important Aspergillus and Penicillium species. In: Sariaslani S, Gadd GM (eds) Advances in applied microbiology. Academic Press, San Diego, pp 199–249 Houbraken J, Wang L, Lee HB, Frisvad JC (2016) New sections in Penicillium containing novel species producing patulin, pyripyropens or other bioactive compounds. Persoonia 36:299–314 Hu DM, Cai L, Chen H, Bahkali AH, Hyde KD (2010) Four new freshwater fungi associated with submerged wood from Southwest Asia. Sydowia 62:191–203 Huang SK, Maharachchikumbura SS, Jeewon R, Bhat DJ, Chomnunti P, Hyde KD, Lumyong S (2018) Morphological and molecular taxonomy of Jahnula dianchia sp. nov. (Jahnulales) from submerged wood in Dianchi Lake, Yunnan China. Mycol Prog 17:547–555 Huanraluek N, Thambugala KM, Wang Y, Hyde KD (2018) Introducing Massarioramusculicola, a novel genus in Massariaceae. Phytotaxa 371:17–24 Hughes SJ, Illman WI (1974a) Sporidesmium foliculatum. Fungi Canadenses 60:1–2 Hughes SJ, Illman WI (1974b) Sporidesmium brachypus. Fungi Canadenses 57:1–2 Huhndorf SM, Miller AN, Fernández FA (2004) Molecular systematics of the Sordariales: the order and the family Lasiosphaeriaceae redefined. Mycologia 96:368–387 Hyde KD (1989) Ecology of tropical marine fungi. Hydrobiologia 178:199–208 Hyde KD (1991) A new amphisphaeriaceous fungus from intertidal fronds of Nypa fruticans. Trans Mycol Soc Jpn 32:265–267 Hyde KD (1992) Intertidal mangrove fungi from the west coast of Mexico, including one new genus and two new species. Mycol Res 96:25–30 123 Fungal Diversity (2019) 96:1–242 Hyde KD (1993) Tropical Australian freshwater fungi. V. Bombardia sp., Jahnula australiensis sp. nov., Savoryella aquatica sp. nov. and S. lignicola sp. nov. Aust Syst Bot 6:161–167 Hyde KD (1995a) Fungi from palms. 17. The genus Fasciatispora, with notes on Amphisphaerella. Nova Hed 61:249–268 Hyde KD (1995b) Tropical Australasian fungi. IX. Vaginatispora aquatica gen. et sp. nov. Nova Hed 61:233–241 Hyde KD, Alias SA (2000) Biodiversity and distribution of fungi associated with decomposing Nypa fruticans. Biodivers Conserv 9:393–402 Hyde KD, Goh TK (1998) Fungi on submerged wood in Lake Barrine, North Queensland, Australia. Mycol Res 102:739–749 Hyde KD, Sarma VV (2006) Biodiversity and ecological observations on filamentous fungi of mangrove palm Nypa fruticans Wurumb (Liliopsida – Arecales) along the Tutong River, Brunei. Indian J Geo-Mar Sci 4:297–307 Hyde KD, Wong SW (1999) Tropical Australian freshwater Fungi. XV. The ascomycete genus Jahnula, with five new species and one new combination. Nova Hed 68:489–509 Hyde KD, Vrijmoed L, Chinnaraj S, Jones EBG (1992) Massarina armatispora sp. nov., a new intertidal ascomycete from mangroves. JAMA 35:325–328 Hyde KD, Jones EBG, Liu JK, Ariyawansa H, Boehm E, Boonmee S, Braun U, Chomnunti P, Crous PW, Dai DQ, Diederich P, Dissanayake A, Doilom M, Doveri F, Hongsanan S, Jayawardena R, Lawrey JD, Li YM, Liu YX, Lücking R, Monka J, Muggia L, Nelsen MP, Pang KL, Phookamsak R, Senanayake IC, Shearer CA, Suetrong S, Tanaka K, Thambugala KM, Wijayawardene NN, Wikee S, Wu HX, Zhang Y, Begoña AH, Alias SA, Aptroot A, Bahkali AH, Bezerra JL, Bhat DJ, Camporesi E, Chukea E, Gueidan C, Hawksworth DL, Hirayama K, Hoog SD, Kang JK, Knudsen K, Li WJ, Li XH, Liu ZY, Mapook A, Mckenzie EHC, Miller AN, Mortimer PE, Phillips AJL, Raja HA, Scheuer C, Schumm F, Taylor JE, Tian Q, Tibpromma S, Wanasinghe DN, Wang Y, Xu JC, Yacharoen S, Yan JY, Zang M (2013) Families of Dothideomycetes. Fungal Divers 63:1–313 Hyde KD, Hongsanan S, Jeewon R, Bhat DJ, McKenzie EHC, Jones EBG, Phookamsak R, Ariyawansa HA, Boonmee S, Zhao Q, Abdel-Aziz FA, Abdel-Wahab MA, Banmai S, Chomnunti P, Cui BK, Daranagama DA, Das K, Dayarathne MC, de Silva NI, Dissanayake AJ, Doilom M, Ekanayaka AH, Gibertoni TB, Góes-Neto A, Huang SK, Jayasiri SC, Jayawardena RS, Konta S, Lee HB, Li WJ, Lin CG, Liu JK, Lu YZ, Luo ZL, Manawasinghe IS, Manimohan P, Mapook A, Niskanen T, Norphanphoun C, Papizadeh M, Perera RH, Phukhamsakda C, Richter C, de Santiago ALCMA, Drechsler-Santos ER, Senanayake IC, Tanaka K, Tennakoon TMDS, Thambugala KM, Tian Q, Tibpromma S, Thongbai B, Vizzini A, Wanasinghe DN, Wijayawardene NN, Wu HX, Yang J, Zeng XY, Zhang H, Zhang JF, Bulgakov TS, Camporesi E, Bahkali AH, Amoozegar AM, Araujo-Neta LS, Ammirati JF, Baghela A, Bhatt RP, Bojantchev S, Buyck B, da Silva GA, de Lima CLF, de Oliveira RJV, de Souza CAF, Dai YC, Dima B, Duong TT, Ercole E, Mafalda-Freire F, Ghosh A, Hashimoto A, Kamolhan S, Kang JC, Karunarathna SC, Kirk PM, Kytövuori I, Lantieri A, Liimatainen K, Liu ZY, Liu XZ, Lücking R, Medardi G, Mortimer PE, Nguyen TTT, Promputtha I, Raj KNA, Reck MA, Lumyong S, Shahzadeh-Fazeli SA, Stadler M, Soudi MR, Su HY, Takahashi T, Tangthirasunun N, Uniyal P, Wang Y, Wen TC, Xu JC, Zhang ZK, Zhao YC, Zhou JZ, Zhu L (2016) Fungal diversity notes 367–490: taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 80:1–270 Hyde KD, Maharachchikumbura SSN, Hongsanan S, Samarakoon MC, Lücking R, Pem D, Harishchandra D, Jeewon R, Zhao RL, Fungal Diversity (2019) 96:1–242 Xu JC (2017a) The ranking of fungi: a tribute to David L. Hawksworth on his 70th birthday. Fungal Divers 84:1–23 Hyde KD, Norphanphoun C, Abreu VP, Bazzicalupo A, Chethana KWT, Clericuzio M, Dayarathne MC, Dissanayake AJ, Ekanayaka AH, He MQ, Hongsanan S, Huang SK, Jayasiri SC, Jayawardena RS, Karunarathna A, Konta S, Kušan I, Lee H, Li J, Lin CG, Liu NG, Lu YZ, Luo ZL, Manawasinghe IS, Mapook A, Perera RH, Phookamsak R, Phukhamsakda C, Siedlecki I, Soares AM, Tennakoon DS, Tian Q, Tibpromma S, Wanasinghe DN, Xiao YP, Yang J, Zeng XY, Abdel-Aziz FA, Li WJ, Senanayake IC, Shang QJ, Daranagama DA, De Silva NI, Thambugala KM, Abdel-Wahab MA, Bahkali AH, Berbee ML, Boonmee S, Bhat DJ, Bulgakov TS, Buyck B, Camporesi E, Castañeda-Ruiz RF, Chomnunti P, Doilom M, Dovana F, Gibertoni TB, Jadan M, Jeewon R, Jones GEB, Kang JC, Karunarathna SC, Lim YW, Liu JK, Liu ZY, Plautz HL Jr, Lumyong S, Maharachchikumbura SSN, Matočec N, Mckenzie EHC, Meśič A, Miller D, Pawłowska J, Pereira OL, Promputtha I, Romero AI, Ryvarden L, Su HY, Suetrong S, Tkalčec Z, Vizzini A, Wen TC, Wisitrassameewong K, Wrzosek M, Xu JC, Zhao Q, Zhao RL, Mortimer PE (2017b) Fungal diversity notes 603–708: taxonomic and phylogenetic notes on genera and species. Fungal Divers 87:1–235 Hyde KD, Chaiwan N, Norphanphoun C, Boonmee S, Camporesi E, Chethana KWT, Dayarathne MC, de Silva NI, Dissanayake AJ, Ekanayaka AH, Hongsanan S, Huang SK, Jayasiri SC, Jayawardena RS, Jiang HB, Karunarathna A, Lin CG, Liu JK, Liu NG, Lu YZ, Luo ZL, Maharachchimbura SSN, Manawasinghe IS, Pem D, Perera RH, Phukhamsakda C, Samarakoon MC, Senwanna C, Shang Q, Tennakoon DS, Thambugala KM, Tibpromma S, Wanasinghe DN, Xiao YP, Yang J, Zeng XY, Zhang JF, Zhang SN, Bulgakov TS, Bhat DJ, Cheewangkoon R, Goh TK, Jones EBG, Kang JC, Jeewon R, Liu ZY, Lumyong S, Kuo CH, McKenzie EHC, Wen TC, Yan JY, Zhao Q (2018a) Mycosphere notes 169–224. Mycosphere 9:271–430 Hyde KD, Norphanphoun C, Chen J, Dissanayake AJ, Doilom M, Hongsanan S, Jayawardena RS, Jeewon R, Perera RH, Thongbai B, Wanasinghe DN (2018b) Thailand’s amazing diversity: up to 96% of fungi in northern Thailand may be novel. Fungal Divers 93:215–239 Hywel-Jones NL (1994) Cordyceps khaoyaiensis and C. pseudomilitaris, two new pathogens of lepidopteran larvae from Thailand. Mycol Res 98:939–942 Inderbitzin P, Landvik S, Abdel-Wahab MA, Berbee ML (2001) Aliquandostipitaceae, a new Crousfamily for two new tropical ascomycetes with unusually wide hyphae and dimorphic ascomata. Am J Bot 88:52–61 Index Fungorum (2019) Index Fungorum. http://www.indexfun gorum.org/ Irinyi L, Serena C, Garcia-Hermoso D, Arabatzis M, Desnos-Ollivier M, Vu D, Cardinali G, Arthur I, Normand AC, Giraldo A, da Cunha KC, Sandoval-Denis M, Hendrickx M, Nishikaku AS, Melo ASD, Merseguel KB, Khan A, Parente Rocha JA, Sampaio P, Briones MRD, Ferreira RCE, Muniz MD, Castanon-Olivares LR, Estrada-Barcenas D, Cassagne C, Mary C, Duan SY, Kong FR, Sun AY, Zeng XY, Zhao ZT, Gantois N, Botterel F, Robbertse B, Schoch C, Gams W, Ellis D, Halliday C, Chen S, Sorrell TC, Piarroux R, Colombo AL, Pais C, de Hoog S, Zancope-Oliveira RM, Taylor ML, Toriello C, Soares CMD, Delhaes L, Stubbe D, Dromer F, Ranque S, Guarro J, Cano-Lira JF, Robert V, Velegraki A, Meyer W (2015) International Society of Human and Animal Mycology (ISHAM)-ITS reference DNA barcoding database-the quality controlled standard tool for routine identification of human and animal pathogenic fungi. Med Mycol 53:313–337 229 Jaklitsch WM, Voglmayr H (2017) Three former taxa of Cucurbitaria and considerations on Petrakia in the Melanommataceae. Sydowia 69:81–95 Jaklitsch WM, Gardiennet A, Voglmayr H (2016) Resolution of morphology-based taxonomic delusions: Acrocordiella, Basiseptospora, Blogiascospora, Clypeosphaeria, Hymenopleella, Lepteutypa, Pseudapiospora, Requienella, Seiridium and Strickeria. Persoonia 37:82–105 Jaklitsch VM, Checa J, Blanco MN, Olariaga I, Tello S, Voglmayr H (2017) A preliminary account of the Cucurbitariaceae. Stud Mycol 90:71–118 Jaklitsch WM, Checa J, Blanco MN, Olariaga I, Tello S, Voglmayr H (2018) A preliminary account of the Cucurbitariaceae. Stud Mycol 90:71–118 Jayasiri SC, Hyde KD, Jones EBG, Jeewon R, Ariyawansa HA, Bhat JD, Camporesi E, Kang JC (2017) Taxonomy and multigene phylogenetic evaluation of novel species in Boeremia and Epicoccum with new records of Ascochyta and Didymella (Didymellaceae). Mycosphere 8:1080–1101 Jayasiri SC, Hyde KD, Jones EBG, Peršoh D, Camporesi E, Kang JC (2018) Taxonomic novelties of hysteriform Dothideomycetes. Mycosphere 9:803–837 Jayasiri SC, Hyde KD, Jones EBG, McKenzie EHC, Jeewon R, Phillips AJL, Bhat DJ, Wanasinghe DN, Liu JK, Lu YZ, Kang JC, Xu J, Karunarathna SC (2019) Diversity, morphology and molecular phylogeny of Dothideomycetes on decaying wild seed pods and fruits. Mycosphere 10:1–186 Jayawardena RS, Hyde KD, Chethana KWT, Daranagama DA, Dissanayake AJ, Goonasekara ID, Manawasinghe IS, Mapook A, Jayasiri SC, Karunarathna A, Li CG (2018) Mycosphere Notes 102–168: saprotrophic fungi on Vitis in China, Italy, Russia and Thailand. Mycosphere 9:1–114 Jeewon R, Liew EC, Simpson JA, Hodgkiss IJ, Hyde KD (2003) Phylogenetic significance of morphological characters in the taxonomy of Pestalotiopsis species. Mol Phylogenet Evol 27:372–383 Jeewon R, Hyde KD (2016) Establishing species boundaries and new taxa among fungi: recommendations to resolve taxonomic ambiguities. Mycosphere 7:1669–1677 Ji XH, He SH, Chen JJ, Si J, Wu F, Zhou LW, Vlasák J, Tian XM, Dai YC (2017) Global diversity and phylogeny of Onnia (Hymenochaetaceae) species on gymnosperms. Mycologia 109:27–34 Ji XH, Vlasák J, Tian XM, Dai YC (2018) Three new species of Fomitiporella (Hymenochaetales, Basidiomycota) based on the evidence from morphology and DNA sequence data. MycoKeys 30:3–89 Jiang XM, Li YK, Liang JF, Wu JR (2017) Russula brunneovinacea sp. nov., from northeastern China. Mycotaxon 132:789–797 Johnston PR, Seifert KA, Stone JK, Rossman AY, Marvanová L (2014) Recommendations on generic names competing for use in Leotiomycetes (Ascomycota). IMA Fungus 5:91–120 Jones EBG, Suetrong S, Sakayaroj J, Bahkali AH, Abdel-Wahab MA, Boekhout T, Pang K-L (2015) Classification of marine Ascomycota, Basidiomycota, Blastocladiomycota and Chytridiomycota. Fungal Divers 73:1–72 Jones EBG, Pang KL, Abdel-Wahab MA, Scholz B, Hyde KD, Boekhout T, Ebel R, Rateb ME, Henderson L, Sakayaroj J, Suetrong S, Dayarathne MC, Kumar V, Raghukumar S, Sridhar KR, Bahkali AHA, Gleason FH, Norphanphoun C (2019) An online resource for marine fungi. Fungal Divers. https://doi.org/ 10.1007/s13225-019-00426-5 Joshi Y, Koh YJ, Hur JS (2010) Further Additions to Lichen Genus Buellia De Not. in South Korea. Mycobiology 38:222–224 123 230 Jukes TH, Cantor CR (1969) Evolution of protein molecules. In: Munro HN (ed) Mammalian protein metabolism. Academic Press, New York, pp 21–132 Jülich W (1982) Higher taxa of Basidiomycetes. Bibl Mycol 85:1–485 Justo A, Miettinen O, Floudas D, Ortiz-Santana B, Sjokvist E, Lindner D, Nakasone K, Niemelä T, Larsson K, Ryvarden L, Hibbett DS (2017) A revised family-level classification of the Polyporales (Basidiomycota). Fungal Biol 121:798–824 Kaishian P, Weir A (2018) New species of Prolixandromyces (Laboulbeniales) from South America. Mycologia 110:222–229 Kalb K (1986) Lichenes Neotropici, Fasc. IX, nos. 351–400:1–16. Neumarkt Kalb K, Aptroot A (2018) New lichen species from Brazil and Venezuela. Bryologist 121:56–67 Kalb K, Plata ER, Lücking R, Lumbsch HT (2011) The phylogenetic position of Malmidea, a new genus for the Lecidea piperis- and Lecanora granifera-groups (Lecanorales, Malmideaceae), inferred from nuclear and mitochondrial ribosomal DNA sequences, with special reference to Thai species. Bibl Lichenol 106:143–168 Kang JC, Hyde KD, Kong YC (1999) Studies on Amphisphaeriales: the genera excluded from the Amphisphaeriaceae, Cainiaceae and Clypeosphaeriaceae. Fungal Divers 2:135–151 Karasińsk D (2015) A new species of palm-associated Grammothele (Basidiomycota, Polyporales) from Bolivia. Nova Hedw 101:103–110 Karunarathna A, Papizadeh M, Senanayake IC, Jeewon R, Phookamsak R, Goonasekara ID, Wanasinghe DN, Wijayawardene NN, Amoozegar MA, Fazeli SS, Camporesi E (2017) Novel fungal species of Phaeosphaeriaceae with an asexual/sexual morph connection. Mycosphere 8:1818–1834 Kashiwadani H, Kalb K (1993) The genus Ramalina in Brazil. Lichenologist 25:1–31 Kawamura S (1929) On some new Japanese fungi. Jap J Bot 4:291–302 Keirle MR, Hemmes DE, Desjardin DE (2004) Agaricales of the Hawaiian Islands. 8. Agaricaceae: Coprinus and Podaxis; Psathyrellaceae: Coprinopsis, Coprinellus and Parasola. Fungal Divers 15:33–124 Kelemu S, White JF Jr, Muñoz F, Takayama Y (2001) An endophyte of the tropical forage grass Brachiaria brizantha: isolating, identifying, and characterizing the fungus, and determining its antimycotic properties. Can J Microbiol 47:55–62 Kendrick B (2000) The fifth kingdom. Hackett Publishing, Indianapolis Kepler RM, Luangsa-Ard JJ, Hywel-Jones NL, Quandt CA, Sung GH, Rehner SA, Aime MC, Henkel TW, Sanjuan T, Zare R, Chen M (2017) A phylogenetically-based nomenclature for Cordycipitaceae (Hypocreales). IMA Fungus 8:335–353 Kimura M (1980) A simple method for estimating evolutionary rate of base substitutions through comparative studies of nucleotide sequences. J Mol Evol 16:111–120 Knight SJ, Klaere S, Morrison-Whittle P, Goddard MR (2018) Fungal diversity during fermentation correlates with thiol concentration in wine. Aust J Grape Wine Res 24:105–112 Kirk PM (1981) New or interesting microfungi II. Dematiaceous Hyphomycetes from Ester Common, Surrey. Trans Br Mycol Soc 77:279–297 Kirk PM, Cannon PF, David JC, Stalpers JA (2001) Ainsworth & Bisby’s dictionary of the fungi, 9th edn. CABI Publishing, UK Kirk PM, Cannon PF, Minter DW, Stalpers JA (2008) Ainsworth & Bisby’s dictionary of the fungi, 10th edn. CABI Publishing, Wallingford 123 Fungal Diversity (2019) 96:1–242 Kirschstein W (1936) Beiträge zur Kenntnis der Ascomyceten und ihrer Nebenformen besonders aus der Mark Brandenburg und dem Bayerischen Walde. Ann Mycol 34:180–210 Klaubauf S, Tharreau D, Fournier E, Groenewald JZ, Crous PW, de Vries RP, Lebrun MH (2014) Resolving the polyphyletic nature of Pyricularia (Pyriculariaceae). Stud Mycol 79:85–120 Konta S, Hongsanan S, Eungwanichayapant PD, Liu JK, Jeewon R, Hyde KD, Maharachchikumbura SSN, Boonmee S (2017) Leptosporella (Leptosporellaceae fam. nov.) and Linocarpon and Neolinocarpon (Linocarpaceae fam. nov.) are accommodated in Chaetosphaeriales. Mycosphere 8:1943–1974 Körber GW (1855) Systema Lichenum Germaniae (I–XXXIV). Die Flechten Deutschlands mikroscopisch gepruft, kritisch gesichtet, charakteristisch beschrieben und systematisch geordnet. Verlag Trewendt & Granier, Breslau, pp 1–458 Korf RP, Zhuang WY (1984) The ellipsoid-spored species of Pulvinula (Pezizales). Mycotaxon 20:607–616 Korf RP, Zhuang WY (1991) A preliminary discomycete flora of Macaronesia: part 16, Otideaceae, Scutellinioideae. Mycotaxon 40:79–106 Koukol O, Delgado G, Hofmann TA, Piepenbring M (2018) Panama, a hot spot for Hermatomyces (Hermatomycetaceae, Pleosporales) with five new species, and a critical synopsis of the genus. IMA Fungus 9:107–141 Kreisel H (1969) Grundzüge eines natürlichen Systems der Pilze. Cramer, Lehre Kristiansen JVBR (1999) A new species of Boubovia (Pezizales). Persoonia 17(2):265–271 Kristiansen R (2016) The genus Boubovia (Velen.) Svrček in Norway. Agarica 37:51–65 Kumar S, Stecher G, Tamura K (2016) MEGA7: Molecular Evolutionary Genetics Analysis version 7.0 for bigger datasets. Mol Biol Evol 33:1870–1874 Kušan I, Matočec N, Jadan M, Tkalčec Z, Mešić A (2018) An overview of the genus Coprotus (Pezizales, Ascomycota) with notes on the type species and description of C. epithecioides sp. nov. MycoKeys 29:15–47 Langer G (1994) Die GattungBotryobasidium Donk (Corticiaceae, Basidiomycetes). Bibl Mycol 158:1–459 Larsen MJ, Cobb-Poulle LA (1990) Phellinus (Hymenochaetaceae). A survey of the world taxa. Syn Fungorum 3:1–206 Larsson KH (2007) Re-thinking the classification of corticioid fungi. Mycol Res 111:1040–1063 Larsson E, Larsson KH (2003) Phylogenetic relationships of russuloid basidiomycetes with emphasis on aphyllophoralean taxa. Mycologia 95:1037–1065 Larsson KH, Parmasto E, Fischer M, Langer E, Nakasone KK, Redhead SA (2007) Hymenochaetales: a phylogeny of the hymenochaetoid clade. Mycologia 98:926–936 Lawrey JD, Zimmermann E, Sikaroodi M, Diederich P (2016) Phylogenetic diversity of bulbil-forming lichenicolous fungi in Cantharellales including a new genus and species. Bryology 119:341–349 Léger JC (1998) Le genre Hymenochaete Léveillé. Bibl Mycol 171:1–319 Letcher PM, McGee PA, Powell MJ (2004) Zoosporic fungi from soils of New South Wales. Can J Bot 82:1490–1500 Letcher PM, Powell MJ, Churchill PF, Chambers JG (2006) Ultrastructural and molecular phylogenetic delineation of a new order, the Rhizophydiales (Chytridiomycota). Mycol Res 110:898–915 Letcher PM, Powell MJ, Davis WJ (2015) A new family and four new genera in Rhizophydiales (Chytridiomycota). Mycologia 107:808–830 Léveillé JH (1845) Champignons exotiques. Ann Sci Nat Bot 3:38–71 Fungal Diversity (2019) 96:1–242 Li GJ (2014) Taxonomy of Russula from China. PhD dissertation, Institute of Microbiology, Chinese Academy of Sciences & University of Chinese Academy of Sciences, Beijing Li GJ, Zhao Q, Zhao D, Yue SF, Li SF, Wen HA, Liu XZ (2013) Russula atroaeruginea and R. sichuanensis spp. nov. from southwest China. Mycotaxon 124:173–188 Li W, Liu J, Bhat DJ, Camporesi E, Xu J, Hyde KD (2014) Introducing the novel species, Dothiorella symphoricarposicola, from snowberry in Italy. Cryptogam, Mycol 35:257–271 Li GJ, Li SF, Zhao D, Wen HA (2015) Recent research progress of Russula (Russulales, Agaricomycetes): a review. Mycosystema 34:821–848 Li GJ, Hyde KD, Zhao RN, Hongsanan S, Abdel-Aziz FA, AbdelWahab MA, Alvarado P, Alves-Silva G, Ammirati JF, Ariyawansa HA, Baghela A, Bahkali AH, Beug M, Bhat DJ, Bojantchev D, Boonpratuang T, Bulgakov TS, Camporesi E, Boro MC, Ceska O, Chakraborty D, Chen JJ, Chethana KWT, Chomnunti P, Consiglio G, Cui BK, Dai DQ, Dai YC, Daranagama DA, Das K, Dayarathne MC, Crop ED, De Oliveira RJV, de Souza CAF, de Souza JI, Dentinger BTM, Dissanayake AJ, Doilom M, Drechsler-Santos ER, Ghobad-Nejhad M, Gilmore SP, Góes-Neto A, Gorczak M, Haitjema GH, Hapuarachchi KK, Hashimoto A, He MQ, Henske JK, Hirayama K, Iribarren MJ, Jayasiri SC, Jayawardena RS, Jeon SJ, Jerônimo GH, Jesus AL, Jones EBG, Kang JC, Karunarathna SC, Kirk PM, Konta S, Kuhnert E, Langer E, Lee HS, Lee HB, Li WJ, Li XH, Liimatainen K, Lima DX, Lin CG, Liu JK, Liu XZ, Liu ZY, Luangsa-ard JJ, Lücking R, Lumbsch HT, Lumyong S, Leaño EM, Marano AV, Matsumura M, McKenzie EHC, Mongkolsamrit S, Mortimer PE, Nguyen TTT, Niskanen T, Norphanphoun C, O’Malley MA, Parnmen S, Pawłowska J, Perera RH, Phookamsak R, Phukhamsakda C, Pires-Zottarelli CLA, Raspé O, Reck MA, Rocha SCO, de Santiago ALCMA, Senanayake IC, Setti L, Shang QJ, Singh SK, Sir EB, Solomon KV, Song J, Srikitikulchai P, Stadler M, Suetrong S, Takahashi H, Takahashi T, Tanaka K, Tang LP, Thambugala KM, Thanakitpipattana D, Theodorou MK, Thongbai B, Thummarukcharoen T, Tian Q, Tibpromma S, Verbeken A, Vizzini A, Vlasák J, Voigt K, Wanasinghe DN, Wang Y, Weerakoon G, Wen HA, Wen TC, Wijayawardene NN, Wongkanoun S, Wrzosek M, Xiao YP, Xu JC, Yan JY, Yang J, Yang SD, Hu Y, Zhang JF, Zhao J, Zhou LW, Peršoh D, Phillips AJL, Maharachchikumbura SSN (2016) Fungal diversity notes 253–366: taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 78:1–237 Li J-F, Phookamsak R, Jeewon R, Bhat DJ, Mapook A, Camporesi E, Shang Q-J, Chukeatirote E, Bahkali AH, Hyde KD (2017) Molecular taxonomy and morphological characterization reveal new species and new host records of Torula species (Torulaceae, Pleosporales). Mycol Prog 16:447–461 Li GJ, Zhang CL, Zhao RL, Lin FC (2018a) Two new species of Russula from Northeast China. Mycosphere 9:431–443 Li GJ, Zhang CL, Zhao RL, Lin FC (2018b) Hypogeous gasteroid Lactarius sulphosmus sp. nov. and agaricoid Russula vinosobrunneola sp. nov. (Russulaceae) from China. Mycosphere 9:838–858 Li GJ, Zhao RL, Zhang CL, Lin FC (2018c) A preliminary DNA barcode selection for the genus Russula (Russulales, Basidiomycota). Mycology. https://doi.org/10.1080/21501203.2018. 1500400 Liew EC, Aptroot A, Hyde KD (2002) An evaluation of the monophyly of Massarina based on ribosomal DNA sequences. Mycologia 94:803–813 Lindemann U, Vega M, Alvarado P (2015) Revision der Gattung Kotlabaea: K. deformis, K. delectans und K. benkertii. Z Mykol 81:373–402 Link (1826) Abh. dt. Akad. Wiss. Berlin 1824:165 231 Link JHF (1833) Handbuch zur Erkennung der Nutzbarsten und am Haufigsten Vorkommenden Gewachse, 3. Spenerschen Buchhandlung, Berlin, p 347348 Liu XY, Udayanga D, Luo ZL, Chen LJ, Zhou DQ, Su HY, Hyde KD (2015) Backbone tree for Chaetothyriales with four new species of Minimelanolocus from aquatic habitats. Fungal Biol 119:1046–1062 Liu JK, Yang J, Maharachchikumbura SSN, McKenzie EHC, Jones EBG, Hyde KD, Liu ZY (2016) Novel chaetosphaeriaceous hyphomycetes from aquatic habitats. Mycol Prog 15:1157–1167 Liu JK, Hyde KD, Jeewon R, Phillips AJ, Maharachchikumbura SS, Ryberg M, Liu ZY, Zhao Q (2017) Ranking higher taxa using divergence times: a case study in Dothideomycetes. Fungal Divers 84:75–99 Liu JK, Lu YZ, Cheewangkoon R, To-Anun C (2018) Phylogeny and morphology of Helicotubeufia gen. nov., with three new species in Tubeufiaceae from aquatic habitats. Mycosphere 9:495–509 Locquin M (1984) Mycologie Generale et Structurale. Masson, Paris Lofgren LA, LeBlanc NR, Certano AK, Nachtigall J, LaBine KM, Riddle J, Broz K, Dong Y, Bethan B, Kafer CW, Kistler HC (2018) Fusarium graminearum: pathogen or endophyte of North American grasses? New Phytol 217:1203–1212 Lombard L, Van der Merwe N, Groenewald J, Crous PW (2015) Generic concepts in Nectriaceae. Stud Mycol 80:189–245 Longcore JE (1996) Chytridiomycete taxonomy since 1960. Mycotaxon 60:149–174 Lopes UP, Zambolim L, Lopes UN, Pereira OL, Costa H (2010) First report of Pilidium concavum causing tan-brown rot in strawberry fruits in Brazil. Plant Pathol 59:1171–1172 Lotsy JP (1907) Vorträge über botanische Stammesgeschichte [Lectures on botanical phylogeny, held at the Imperial University of Leiden]. Gustav Fischer, Jena Lowe JL (1964) The genera Grammothele and Porogramme. Pap Mich Acad Sci 49:27–40 Lu YZ (2018) Multi-gene phylogeny and morphology reappraisal of Tubeufiales. PhD Thesis, Mae Fah Luang University, Chiang Rai, Thailand Lu YZ, Boonmee S, Bhat DJ, Hyde KD, Kang JC (2017) Helicosporium luteosporum sp. nov. and Acanthohelicospora aurea (Tubeufiaceae, Tubeufiales) from terrestrial habitats. Phytotaxa 319:241–253 Lu YZ, Boonmee S, Liu JK, Hyde KD, McKenzie EH, Eungwanichayapant PD, Kang JC (2018a) Multi-gene phylogenetic analyses reveals Neohelicosporium gen. nov. and five new species of helicosporous hyphomycetes from aquatic habitats. Mycol Prog 17:631–646 Lu YZ, Liu KJ, Hyde KD, Bhat DJ, Xiao YP, Tian Q, Wen TC, Boonmee S, Kang JC (2016) Brunneodinemasporium jonesii and Tainosphaeria jonesii spp. nov. (Chaetosphaeriaceae, Chaetosphaeriales) from southern China. Mycosphere 7:1322–1331 Lu YZ, Liu JKJ, Hyde KD, Jeewon R, Kang JC, Fan C, Boonmee S, Bhat DJ, Luo ZL, Lin CG, Eungwanichayapant PD (2018b) A taxonomic reassessment of Tubeufiales based on multi-locus phylogeny and morphology. Fungal Divers 92:131–344 Lücking R, Hodkinson BP, Leavitt SD (2017) The 2016 classification of lichenized fungi in the Ascomycota and Basidiomycota— approaching on thousand genera. Bryologist 119:361–416 Lumbsch HT, Huhndorf SM (2010) Myconet Volume 14. Part One. Outline of Ascomycota—2009. Part Two. Notes on Ascomycete Systematics. Nos. 4751–5113. Fieldiana Life Earth Sci 1:1–64 Lumbsch HT, Schmitt I, Lindemuth R, Miller A, Mangold A, Fernandez F, Huhndorf S (2005) Performance of four ribosomal DNA regions to infer higher-level phylogenetic relationships of non-operculate Euascomycetes (Leotiomyceta). Mol Phylogenet Evol 34:512–524 123 232 Lumbsch HT, Kraichak E, Parnmen S, Rivas Plata E, Aptroot A, Cáceres MES, Ertz S, Feuerstein SC, Mercado-Dı́az JA, Staiger B, Van den Broeck D, Lücking R (2014) New higher taxa in the lichen family Graphidaceae (lichenized Ascomycota: Ostropales) based on a three-gene skeleton phylogeny. Phytotaxa 189:39–51 Luo ZL, Bao DF, Bhat DJ, Yang J, Chai HM, Li SH, Bahkali AH, Su HY, Hyde KD (2016) Sporoschisma from submerged wood in Yunnan, China. Mycol Prog 15:1145–1155 Luo ZL, Bhat DJ, Jeewon R, Boonmee S, Bao DF, Zhao YC, Chai HM, Su HY, Su XJ, Hyde KD (2017) Molecular phylogeny and morphological characterization of asexual fungi (Tubeufiaceae) from freshwater habitats in Yunnan, China. Cryptogam Mycol 38:27–54 Luo ZL, Hyde KD, Liu JK, Bhat DJ, Bao DF, Li WL, Su HY (2018) Lignicolous freshwater fungi from China II: novel Distoseptispora (Distoseptisporaceae) species from northwestern Yunnan Province and a suggested unified method for studying lignicolous freshwater fungi. Mycosphere 9:444–461 Luo ZL, Hyde KD, Liu JK, Maharachchikumbura SSN, Jeewon R, Bao DF, Bhat DJ, Lin CG, Li WL, Yang J, Liu NG, Lu YZ, Jayawardena RS, Li JF, Su HY (2019) Freshwater Sordariomycetes. Fungal Divers (in press) Luttrell ES (1955) The ascostromatic ascomycetes. Mycologia 47:511–532 Magyar D, Shoemaker RA, Bobvos J, Crous PW, Groenewald JZ (2011) Pyrigemmula, anovel hyphomycete genus on grapevine and tree bark. Mycol Prog 10:307–314 Maharachchikumbura SS, Guo LD, Chukeatirote E, Bahkali AH, Hyde KD (2011) Pestalotiopsis—morphology, phylogeny, biochemistry and diversity. Fungal Divers 50:167–187 Maharachchikumbura SSN, Guo LD, Cai L, Chukeatirote E, Wu WP, Sun X, Crous PW, Bhat DJ, McKenzie EHC, Bahkali AH, Hyde KD (2012) A multi-locus backbone tree for Pestalotiopsis, with a polyphasic characterization of 14 new species. Fungal Divers 56:95–129 Maharachchikumbura SSN, Chukeatirote E, Guo LD, Crous PW, Mckenzie EHC, Hyde KD (2013) Pestalotiopsis species associated with Camellia sinensis (tea). Mycotaxon 123:47–61 Maharachchikumbura SS, Hyde KD, Groenewald JZ, Xu J, Crous PW (2014) Pestalotiopsis revisited. Stud Mycol 79:121–186 Maharachchikumbura SSN, Hyde KD, Jones EBG, McKenzie EHC, Huang S-K, Abdel-Wahab MA, Daranagama DA, Dayarathne M, D’souza MJ, Goonasekara ID, Hongsanan S, Jayawardena RS, Kirk PM, Konta S, Liu JK, Liu ZY, Norphanphoun C, Pang KL, Perera RH, Senanayake IC, Shang QJ, Shenoy BD, Xiao YP, Bahkali AH, Kang JC, Somrothipol S, Suetrong S, Wen TC, Xu JC (2015) Towards a natural classification and backbone tree for Sordariomycetes. Fungal Divers 72:199–301 Maharachchikumbura SN, Hyde KD, Jones EBG, McKenzie EHC, Jayarama B, Dayarathne M, Huang SK, Norphanphoun C, Senanayake IC, Perera RH, Shang Q, Xiao Y, Dsouza MJ, Hongsanan S, Jayawardena RS, Daranagama DA, Konta S, Goonasekara ID, Zhuang W-Y, Jeewon R, Phillips AJL, AbdelWahab MA, Al-Sadi AM, Bahkali AH, Boonmee S, Boonyuen N, Cheewangkoon R, Dissanayake AJ, Kang J, Li Q-R, Liu JK, Li XZ, Liu Z-Y, Luangsaard JJ, Pang KL, Phookamsak R, Promputtha I, Suetrong S, Stadler M, Wen T, Wijayawardene NN (2016) Families of Sordariomycetes. Fungal Divers 79:1–317 Mains EB (1959) Cordyceps species. Bull Torrey Bot Club 86:46–58 Mapook A, Boonmee S, Ariyawansa H, Tibpromma S, Camporesi E, Jones EBG, Bahkali AH, Hyde KD (2016) Taxonomic and phylogenetic placement of Nodulosphaeria. Mycol Prog 15:34 Marin-Felix Y, Groenewald JZ, Cai L, Chen Q, Marincowitz S, Barnes I, Bensch K, Braun U, Camporesi E, Damm U, de Beer 123 Fungal Diversity (2019) 96:1–242 ZW, Dissanayake A, Edwards J, Giraldo A, Hernández-Restrepo M, Hyde KD, Jayawardena RS, Lombard L, Luangsa-ard J, McTaggart AR, Rossman AY, Sandoval-Denis M, Shen M, Shivas RG, Tan YP, van der Linde EJ, Wingfield MJ, Wood AR, Zhang JQ, Zhang Y, Crous PW (2017) Genera of phytopathogenic fungi: GOPHY 1. Stud Mycol 86:99–216 Mason EW (1941) Annotated account of fungi received at the Imperial Mycological Institute. Mycol Pap 5:101–144 Mason-Gamer R, Kellog E (1996) Testing for phylogenetic conflict among molecular data sets in the tribe Triticeae (Gramineae). Syst Biol 45:524–545 Matheny PB, Moreau P-A, Vizzini A, Harrower E, Haan AD, Contu M, Curti M (2015) Crassisporium and Romagnesiella: two new genera of dark-spored Agaricales. Syst Biodivers 13:28–41 McCarthy PM, Elix JA (1996) Myeloconis, a new genus of pyrenocarpous lichens from the tropics. Lichenologist 28:401–414 McKenzie EHC (1995) Dematiaceous hyphomycetes on Pandanaceae 5. Sporidesmium sensu lato. Mycotaxon 56:9–29 Mehl JWM, Slippers B, Roux J, Wingfield MJ (2014) Botryosphaeriaceae associated with die-back of Schizolobium parahyba trees in South Africa and Ecuador. For Pathol 44:396–408 Mehrabi M, Hemmati R, Vasilyeva LN, Trouillas FP (2015) A new species and a new record of Diatrypaceae from Iran. Mycosphere 6:60–68 Mehrabi M, Roghayeh Hemmati, Vasilyeva LN, Trouillas FP (2016) Diatrypella macrospora sp. nov. and new records of diatrypaceous fungi from Iran. Phytotaxa 252:43–55 Mejı́a LC, Castlebury LA, Rossman AY, Sogonov MV, White JF Jr (2011) Asystematic account of the genus Plagiostoma (Gnomoniaceae, Diaporthales) based on morphology, host-associations, and a four-gene phylogeny. Stud Mycol 68:211–235 Mel’nik VA (2012) Phaeoisaria vietnamensis sp. nov. and P. clematidis (hyphomycetes) from Vietnam. Mycosphere 3:957–960 Mendonça CO, Aptroot A, Cáceres MES (2016) Six new species of the lichen genus Pyrenula (Pyrenulaceae) from Northeast Brazil. Phytotaxa 286:169–176 Mešić A, Tkalčec Z (2003) Preliminary checklist of Agaricales from Croatia IV: families Bolbitiaceae, Coprinaceae, Entolomataceae and Pluteaceae. Mycotaxon 87:283–309 Miadlikowska J, Kauff F, Högnabba F, Oliver JC, Molnár K, Fraker E, Gaya E, Hafellner J, Hofstetter V, Gueidan C, Otálora MA (2014) A multigene phylogenetic synthesis for the class Lecanoromycetes (Ascomycota): 1307 fungi representing 1139 infrageneric taxa, 317 genera and 66 families. Mol Phyl Evol 79:132–168 Miettinen O, Larsson KH (2011) Sidera, a new genus in Hymenochaetales with poroid and hydnoid species. Mycol Prog 10:131–141 Miller SL, Buyck B (2002) Molecular phylogeny of the genus Russula in Europe with a comparison of modern infrageneric classifications. Mycol Res 106:259–276 Miller SL, Larsson E, Larsson KH, Verbeken A, Nuytinck J (2006) Perspectives in the new Russulales. Mycologia 98:960–970 Miyake I (1909) Studies on the parasitic fungi of rice in Japan. Bot Mag Tokyo 23:85–97 Mohammadi H (2014) Phaeoacremonium spp. and Botryosphaeriaceae spp. associated with Date Palm (Phoenix dactylifera L.) decline in Iran. J Phytopathol 162(9):575–581 Moncalvo JM, Nilsson RH, Koster B, Dunham SM, Bernauer T, Matheny PB, Porter TM, Margaritescu S, Weiß M, Garnica S, Danell E (2006) The cantharelloid clade: dealing with incongruent gene trees and phylogenetic reconstruction methods. Mycologia 98:937–948 Monod M (1983) Monographie taxonomique des Gnomoniaceae. Beih Sydowia 9:1–315 Fungal Diversity (2019) 96:1–242 Montagne JPFC (1834) Notice sur les plantes cryptogames récemment découvertes en France, contenant aussi l’indication précise des localités de quelques espèces les plus rares de la flore française. Ann Sci Nat Bot 1:295–307 Moral J, Agustı́-Brisach C, Pérez-Rodrı́guez M, Xaviér C, Raya MC, Rhouma A, Trapero A (2017) Identification of fungal species associated with branch dieback of olive and resistance of table cultivars to Neofusicoccum mediterraneum and Botryosphaeria dothidea. Plant Dis 101:306–316 Mugambi GK, Huhndorf SM (2009) Molecular phylogenetics of Pleosporales: Melanommataceae and Lophiostomataceae recircumscribed (Plesporomycetidae, Dothideomycetes, Ascomycota). Stud Mycol 64:103–121 Munk A (1953) The system of the Pyrenomycetes. Dansk Bot Arkiv 15:1–163 Mycology and Microbiology Laboratory, ARS, USDA. https://nt. arsgrin.gov/fungaldatabases/ Nagy LG, Walther G, Házi J, Vágvölgyi C, Papp T (2011) Understanding the evolutionary processes of fungal fruiting bodies: correlated evolution and divergence times in the Psathyrellaceae. Syst Biol 60:303–317 Nagy LG, Házi J, Szappanos B, Kocsubé S, Bálint B, Rákhely G, Vágvölgyi C, Papp T (2012) The evolution of defense mechanisms correlate with the explosive diversification of autodigesting Coprinellus Mushrooms (Agaricales, Fungi). Syst Biol 61:595–607 Nagy LG, Desjardin DE, Vágvölgyi C, Kemp R, Papp T (2013) Phylogenetic analyses of Coprinopsis sections Lanatuli and Atramentarii identify multiple species within morphologically defined taxa. Mycologia 105:112–124 Nannfeldt JA (1932) Studien über die Morphologie und Systematik der nichtlichenisierten inoperculaten Discomyceten. Nova Acta Regiae Soc Sci Ups 8:1–368 Nash TH, Ryan BD, Gries C, Bungartz F (eds) (2004) Lichen Flora of the Greater Sonoran Desert Region, vol 2. Arizona State University, Tempe Nitschke TRJ (1869) Grundlage eines systems der Pyrenomyceten. Verhandlungen des Naturhistorischen Vereins der Preussischen Rheinlande, Westfalens und des Regierungsbezirks Osnabrück 26:70–77 Nusbaumer L, Barbosa MRV, Thomas WW, Alves MV, Loizeau PA, Spichiger R (2015) Flora e vegetação da Reserva Biológica de Pedra Talhada. Biodiversidade da Reserva Biológica de Pedra Talhada (Alagoas, Pernambuco—Brasil). Boissiera 68:59–121 Oberwinkler F (1977) Das neue System der Basidiomyceten. In: Frey W, Hurka H, Oberwinkler F (eds) Beiträge zur Biologie der neideren Pflanzen. Gustav Fischer, Stuttgart, pp 59–104 Olariaga I, Buyck B, Esteve-Raventos F, Hofstetter V, Manjón JL, Moreno G, Salcedo I (2015) Assessing the taxonomic identity of white and orange specimens of Cantharellus: occasional colour variants or independent species? Cryptogam, Mycol 36:287–300 Olariaga I, Moreno G, Manjoni JL, Salcedo I, Hofstetter V, Rodrı́guez D, Buyck B (2016) Cantharellus (Cantharellales, Basidiomycota) revisited in Europe through a multigene phylogeny. Fungal Divers 83:263–292 Osmundson TW, Robert VA, Schoch CL, Baker LJ, Smith A, Robich G, Mizzan L, Garbelotto MM (2013) Filling gaps in biodiversity knowledge for macrofungi: contributions and assessment of an herbarium collection DNA barcode sequencing project. PLoS ONE 8:e62419 Palm ME (1991) Taxonomy and morphology of the synanamorphs Pilidium concavum and Hainesia lythri (coelomycetes). Mycologia 83:787–796 Pang KL, Abdel-Wahab MA, Sivichai S, El-Sharouney HM, Jones EBG (2002) Jahnulales (Dothideomycetes, Ascomycota): a new 233 order of lignicolous freshwater ascomycetes. Mycol Res 106:1031–1042 Parmasto E (1968) Conspectus Ssystematis Corticiacearum. Inst Zool Bot, Tartu Parmasto E (2005) New data on rare species of Hydnochaete and Hymenochaete (Hymenochaetales). Mycotaxon 91:137–163 Peck CH (1878) Colorado fungi (new species). Bot Gaz Crawfordsville 3:34–35 Peintner U, Bougher NL, Castellano MA, Moncalvo JM, Moser MM, Trappe JM, Vilgalys R (2001) Multiple origins of sequestrate fungi related to Cortinarius (Cortinariaceae). Am J Bot 88:2168–2179 Peintner U, Horak E, Moser M, Vilgalys R (2002) Phylogeny of Rozites, Cuphocybe and Rapacea inferred from ITS and LSU rDNA sequences. Mycologia 94:620–629 Perdomo H, Sutton DA, Garcia D, Fothergill AW, Cano J, Gene J, Summerbell RC, Rinaldi MG, Guarro J (2011) Spectrum of clinically relevant Acremonium species in the United States. J Clin Microbiol 49:243–256 Pereira LA, Senha KS, Santos MR, Costa Neto SV (2007) Aspectos Florı́sticos da FLONA do Amapá e sua Importância na Conservação da Biodiversidade. Rev Bras Biociênc 5:693–695 Perry BA, Hansen K, Pfister DH (2007) A phylogenetic overview of the family Pyronemataceae (Ascomycota, Pezizales). Mycol Res 111:549–571 Persoon CH (1794) Neuer Versuch einer systematischen Eintheilung der Schwaümme. N Mag Die Bot Ihrem Ganzen Umfange 1:63–128 Persoon CH (1796) Observationes mycologicae. Apud Petrum Phillippum Wolf, Leipzig Phillips AJL, Alves A, Correia A, Luque J (2005) Two new species of Botryosphaeria with brown, 1-septate ascospores and Dothiorella anamorphs. Mycologia 97:513–529 Phillips AJL, Alves A, Johnston P, Pennycook S, Akulov A, Ramaley A, Crous PW (2008) Resolving the status of darkspored teleomorph genera in the Botryosphaeriaceae. Persoonia 21:29–55 Phillips AJL, Alves A, Abdollahzadeh J, Slippers B, Wingfield MJ, Groenewald JZ, Crous PW (2013) The Botryosphaeriaceae: genera and species known from culture. Stud Mycol 76:51–167 Phillips AJ, Hyde KD, Alves A, Liu JK (2019) Families in Botryosphaeriales: a phylogenetic, morphological and evolutionary perspective. Fungal Divers 94:1–22 Phookamsak R, Liu JK, McKenzie EHC, Manamgoda DS, Ariyawansa H, Thambugala KM, Dai DQ, Camporesi E, Chukeatirote E, Wijayawardene NN, Bahkali AH, Mortimer PE, Xu JC, Hyde KD (2014) Revision of Phaeosphaeriaceae. Fungal Divers 68:159–238 Phookamsak R, Wanasinghe DN, Hongsanan S, Phukhamsakda C, Huang SK, Tennakoon DS, Norphanphoun C, Camporesi E, Bulgakov TS, Promputtha I, Mortimer PE (2017) Towards a natural classification of Ophiobolus and ophiobolus-like taxa; introducing three novel genera Ophiobolopsis, Paraophiobolus and Pseudoophiobolus in Phaeosphaeriaceae (Pleosporales). Fungal Divers 87:299–339 Phookamsak R, Lu YZ, Hyde KD, Jeewon R, Li J, Doilom M, Boonmee S, Promputtha I (2018) Phylogenetic characterization of two novel Kamalomyces species in Tubeufiaceae (Tubeufiales). Mycol Prog 17:647–660 Phookamsak R, Hyde KD, Jeewon R, Bhat DJ, Jones EBG, Maharachchikumbura SSN, Raspé O, Karunarathna SC, Wanasinghe DN, Hongsanan S, Doilom M, Tennakoon DS, Machado AR, Firmino AL,Ghosh A, Karunarathna A, Mešić A, Dutta, AK, Thongbai, B, Devadatha, B, Norphanphoun, C, Senwanna, C, Wei, D, Pem, D, Ackah, FK, Wang GN, Jiang HB, Madrid H, Lee HB, Goonasekara ID, Manawasinghe IS, Kušan Cano J, 123 234 Gené J, Li J, Das K, Acharya K, Raj KNA, Latha KPD, Chethana KWT, He MQ, Dueñas M, Jadan M, Martı́n MP, Samarakoon MC, Dayarathne MC, Raza M, Park MS, Telleria MT, Chaiwan N, Matočec N, de Silva NI, Pereira OL, Singh PN, Manimohan P, Uniyal P, Shang QJ, Bhatt RP, Perera RH, Alvarenga RLM, Nogal-Prata S, Singh SK,Vadthanarat S, Oh SY, Huang SK, Rana S, Konta S, Paloi S, Jayasiri SC, Jeon SJ, Mehmood T, Gibertoni TB, Nguyen TTT, Singh U, Thiyagaraja V, Sarma VV, Dong W, Yu XD, Lu YZ, Lim YW, Chen Y, Tkalčec Z, Zhang ZF, Luo ZL, Daranagama DA, Thambugala KM, Tibpromma S, Camporesi E, Bulgakov T, Dissanayake AJ, Senanayake IC, Dai DQ, Tang LZ, Khan S, Zhang H, Promputtha I, Cai L, Chomnunti P, Zhao RL, Lumyong S, Boonmee S, Wen TC, Mortimer PE, Xu J (2019) Fungal diversity notes 929–1036: taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Divers 95:1–273 Phukhamsakda C, Ariyawansa HA, Phookamsak R, Chomnunti P, Bulgakov TS, Yange JB, Bhat DJ, Bahkalih AH, Hyde KD (2015) Muriphaeosphaeria galatellae gen. et. sp. nov. in Phaeosphaeriaceae (Pleosporales). Phytotaxa 227:55–65 Pilantanapak A, Jones EBG, Eaton RA (2005) Marine fungi on Nypa fruticans in Thailand. Bot Mar 48:365–373 Pine EM, Hibbett DS, Donoghue MJ (1999) Phylogenetic relationships of cantharelloid and clavarioid Homobasidiomycetes based on mitochondrial and nuclear rDNA sequences. Mycologia 91:944–963 Pinnoi A, Jeewon R, Sakayaroj J, Hyde KD, Jones EBG (2007) Berkleasmium crunisia sp. nov. and its phylogenetic affinities to the Pleosporales based on 18S and 28S rDNA sequence analyses. Mycologia 99:378–384 Pitt JI, Hocking AD (2009) Fungi and food spoilage, 3rd edn. Springer, New York Pitt WM, Úrbez-Torres JR, Trouillas FP (2013) Dothiorella vidmadera, a novel species from grapevines in Australia and notes on Spencermartinsia. Fungal Divers 61:209–219 Põldmaa K (2011) Tropical species of Cladobotryum and Hypomyces producing red pigments. Stud Mycol 68:1–34 Prasher IB, Sushma P (2014) Hermatomyces indicus sp. nov. (Hyphomycetes) from India. Nova Hedw 99:551–556 Promputtha I, Lumyong S, Dhanasekaran V, McKenzie EHC, Hyde KD, Jeewon R (2007) A phylogenetic evaluation of whether endophytes become saprotrophs at host senescence. Microb Ecol 53:579–590 Punithalingam E (1980) Plant diseases attributed to Botryodiplodia theobromae Pat. J Cramer, Vaduz Quaedvlieg W, Binder M, Groenewald JZ, Summerell BA, Carnegie AJ, Burgess TI, Crous PW (2014) Introducing the consolidated species concept to resolve species in the Teratosphaeriaceae. Persoonia 33:1–40 Rabenhorst GL (1858) Klotzschii herbarium vivum mycologicum sistens fungorum per totam Germaniam crescentium collectionem perfectam. Editio nova. Centuria VIII. No. 701–800 Rabenhorst L (1864–1868) Flora Europaea algarum aquae dulcis et submarinae, vol 3. Verlag Eduard Kummer, Leipzig Raja HA, Shearer CA (2006) Jahnula species from north and central America, including three new species. Mycologia 98:319–332 Raja HA, Carter A, Platt HW, Shearer CA (2008) Freshwater ascomycetes: Jahnula apiospora (Jahnulales, Dothideomycetes), a new species from Prince Edward Island, Canada. Mycoscience 49:326–328 Raja HA, El-Elimat T, Oberlies NH, Shearer CA, Miller AN, Tanaka K, Hashimoto A, Fournier J (2015) Minutisphaerales (Dothideomycetes, Ascomycota): a new order of freshwater ascomycetes including a new family, Minutisphaeraceae, and two new species from North Carolina, USA. Mycologia 107:845–862 123 Fungal Diversity (2019) 96:1–242 Rakotoarinivo M, Razafitsalama JL, Baker WJ, Dransfield J (2010) Analalava—a palm conservation hotspot in eastern Madagascar. Palms 54:141–151 Ranghoo VM (1998) Phylogeny of freshwater ascomycetes. Dissertation, University of Hong Kong Rao VG, Reddy KA (1981) Two new hyphomycetes. Indian J Bot 4:108–114 Rappaz F (1987) Taxonomie et momenclature des Diatrypacees à asques octospores. Mycol Helv 2:285–648 Raut JK, Suzuki A, Fukiharu T, Shimizu K, Kawamoto S, Tanaka C (2011) Coprinopsis neophlyctidospora sp. nov., a new ammonia fungus from boreal forests in Canada. Mycotaxon 115:227–238 Raut JK, Fukiharu T, Shimizu K, Kawamoto S, Takeshige S, Tanaka C, Yamanaka T, Suzuki A (2015) Coprinopsis novorugosobispora (Basidiomycota, Agricales), an ammonia fungus new to Canada. Mycosphere 6:612–619 Reblova M (1999) Studies in Chaetosphaeria sensu lato III. Umbrinosphaeria gen. nov. and Miyoshiella with Sporidesmium anamorphs. Mycotaxon 71:13–43 Réblová M (2004) Four new species of Chaetosphaeria from New Zealand and redescription of Dictyochaeta fuegiana. Stud Mycol 50:171–186 Réblová M, Winka K (2000) Phylogeny of Chaetosphaeria and its anamorphs based on morphological and molecular data. Mycologia 92:939–954 Réblová M, Winka K (2001) Generic concepts and correlations in ascomycetes based on molecular and morphological data: Lecythothecium duriligni gen. et sp. nov. with a Sporidesmium anamorph, and Ascolacicola austriaca sp. nov. Mycologia 93:478–493 Réblová M, Seifert KA (2008) A new species of Chaetosphaeria with Menispora ciliata and phialophora-like synanamorphs. Fungal Divers 29:101–107 Réblová M, Gams W, Seifert KA (2011) Monilochaetes and allied genera of the Glomerellales, and a reconsideration of families in the Microascales. Stud Mycol 68:163–191 Réblová M, Seifert KA, Fournier J, Štěpánek V (2012) Phylogenetic classification of Pleurothecium and Pleurotheciella gen. nov. and its dactylaria-like anamorph (Sordariomycetes) based on nuclear ribosomal and protein-coding genes. Mycologia 104:1299–1314 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:e63547 Réblová M, Seifert KA, Fournier J, Štěpánek V (2016) Newly recognised lineages of perithecial ascomycetes: the new orders Conioscyphales and Pleurotheciales. Persoonia 37:57–81 Redhead SA, Vilgalys R, Moncalvo JM, Johnson J, Hopple JS Jr (2001) Coprinus Pers. and the disposition of Coprinus species sensu lato. Taxon 50:203–241 Rivas Plata E, Lücking R, Sipman HJM, Mangold A, Kalb K, Lumbsch HT (2010) A world-wide key to the thelotremoid Graphidaceae, excuding the Ocellularia-Myriotrema-Stegobolus clade. Lichenologist 42:187–189 Rivas Plata E, Mason-Gamer R, Ashley M, Lücking R, Lumbsch HT (2012) Molecular phylogeny and systematics of the Ocellulariaclade (Ascomycota: Ostropales: Graphidaceae). Taxon 61:1161–1179 Rogers DP (1935) Notes on the lower basidiomycetes. Univ Iowa Stud Nat Hist 17:1–43 Rossi W (1988) New or interesting Laboulbeniales (Ascomycetes) parasitic on Diptera. Webbia 42:171–178 Rossi W, Bernardi M (2018) New species and new records of Laboulbeniales (Ascomycota) from Thailand. Mycosphere 9:1151–1172 Fungal Diversity (2019) 96:1–242 Rossi W, Leonardi M (2018) New species and new records of Laboulbeniales (Ascomycota) from Sierra Leone. Phytotaxa 358:91–116 Rossi W, Weir A (2007) New Species of Stigmatomyces from various continents. Mycologia 99:139–143 Rossi W, Weir A (2011) New species of Stigmatomyces from Asia. Mycologia 103:131–134 Rossi W, Kotrba M, Triebel D (2005) A new species of Stigmatomyces from Baltic amber, the first fossil record of Laboulbeniomycetes. Mycol Res 109:271–274 Rossi W, Santamarı́a S, Andrade R (2013) Notes on the Laboulbeniales (Ascomycota) parasitic on Diptera from Portugal and other countries. Plant Biosyst 147:730–742 Rossman AY, Aime MC, Farr DF, Castlebury LA, Peterson KR, Leahy R (2004) Thecoelomycetous genera Chaetomella and Pilidium represent a newly discovered lineage of inoperculate discomycetes. Mycol Prog 3:275–290 Rossman AY, Farr DF, Castlebury LA (2007) A review of the phylogeny and biology of the Diaporthales. Mycoscience 48:135–144 Rossman AY, Adams GC, Cannon PF, Castlerbury LA, Crous PW, Gryzenhout M, Jaklitsch WM, Mejia LC, Stoykov D, Udayanga D, Voglmayr H, Walker DM (2015a) Recommendations of generic names in Diaporthales competing for protection or use. IMA Fungus 6:145–154 Rossman A, Crous PW, Hyde KD, Hawksworth DL, Aptroot A, Bezerra JL, Bhat JD, Boehm E, Braun U, Boonmee S, Camporesi E (2015b) Recommended names for pleomorphic genera in Dothideomycetes. IMA Fungus 6:507–523 Ruiz A, Cerdán D (2016) El complejo ‘‘radiata’’ del género Coprinopsis, en las zonas ganaderas de Vinuesa (Soria). Bol Micol FAMCAL 11:23–39 Ryvarden L (1979) Porogramme and related genera. Trans Br Mycol Soc 73:9–19 Ryvarden L (1983) Type studies in the Polyporaceae 14. Species described by N. Patouillard, either alone or with other mycologists. Occas Pap Farlow Herb 18:1–38 Ryvarden L (2015) Studies in Neotropical polypores 40 A note on the genus Grammothele. Synop Fungorum 33:36–42 Saccardo PA (1883) Sylloge Fungorum Omnium hucusque cognitorum, vol 2. Pyrenomycologiae universae continuation et finis, Padova Samson RA, Frisvad JC (2004) Penicillium subgenus Penicillium: new taxonomic schemes and mycotoxins and other extrolites. Centraalbureau voor Schimmelcultures, Utrecht Samson RA, Houbraken J, Thrane U, Frisvad JC, Andersen B (2010) Food and indoor fungi. CBS Laboratory Manual Series 2. Centraalbureau voor Schimmelcultures, Utrecht Samuels GJ, Blackwell M (2001) Pyrenomycetes—fungi with perithecia. In: McLaughlin D, McLaughlin E (eds) The Mycota VII Part A. Systematics and evolution. Springer, Berlin, pp 221–255 Sang XY, Li XD, Wang YW, Fan L (2016) Four new sequestrate species of Russulaceae found in China. Phytotaxa 289:101–117 Saravanakumar D, Lavanya N, Muthumeena K, Raguchander T, Samiyappan R (2009) Fluorescent pseudomonad mixtures mediate disease resistance in rice plants against sheath rot (Sarocladium oryzae) disease. Biocontrol 54:273–286 Sarnari M (1998) Monografia illustrate de genere Russula in Europa, Tomo Primo. AMB, Centro Studi Micologici, Trento Scheidegger C (1993) A revision of the European saxicolous species of the genus Buellia De Not. and formerly included genera. Lichenologist 25:315–364 Schenk A (1858) Algologische Mittheilungen. Verh Phys Med Ges Würzbg 8:235–259 235 Schmull M, Miadlikowska J, Pelzer M, Stocker-Wörgötter E, Hofstetter V, Fraker E, Hodkinson BP, Reeb V, Kukwa M, Lumbsch HT, Kauff F (2011) Phylogenetic affiliations of members of the heterogeneous lichen-forming fungi of the genus Lecidea sensu Zahlbruckner (Lecanoromycetes, Ascomycota). Mycologia 103:983–1003 Schoch CL, Shoemaker RA, Seifert KA, Hambleton S, Spatafora JW, Crous PW (2006) A multigene phylogeny of the Dothideomycetes using four nuclear loci. Mycologia 98:1041–1052 Schoch CL, Crous PW, Groenewald JZ, Boehm EWA, Burgess TI, de Gruyter J, de Hoog GS, Dixon LJ, Grube M, Gueidan C, Harada Y, Hatakeyama S, Hirayama K, Hosoya T, Huhndorf SM, Hyde KD, Jones EBG, Kohlmeyer J, Kruys Å, Li YM, Lücking R, Lumbsch HT, Marvanová L, Mbatchou JS, McVay AH, Miller AN, Mugambi GK, Muggia L, Nelsen MP, Nelson P, Owensby CA, Phillips AJL, Phongpaichit S, Pointing SB, Pujade-Renaud V, Raja HA, Rivas Plata E, Robbertse B, Ruibal C, Sakayaroj J, Sano T, Selbmann L, Shearer CA, Shirouzu T, Slippers B, Suetrong S, Tanaka K, Volkmann-Kohlmeyer B, Wingfield MJ, Wood AR, Woudenberg JHC, Yonezawa H, Zhang Y, Spatafora JW (2009) A class-wide phylogenetic assessment of Dothideomycetes. Stud Mycol 64:1–15 Schröter J (1893) Die Pilze Schlesiens. In Cohn F (ed) KryptogamenFlora von Schlesien 3:31 Seifert KA (1985) A monograph of Stilbella and some allied hyphomycetes. Stud Mycol 27:1–235 Seifert K, Morgan-Jones G, Gams W, Kendrick B (2011) The genera of Hyphomycetes. CBS Biodiversity Series 9. CBS KNAW Fungal Biodiversity Centre, Utrecht Senanayake IC, Maharachchikumbura SSN, Hyde KD, Bhat JD, Jones EBG, McKenzie EHC, Dai DQ, Daranagama DA, Dayarathne MC, Goonasekara ID, Konta S, Li WJ, Shang QJ, Stadler M, Wijayawardene NN, Xiao YP, Norphanphoun C, Li Q, Liu XZ, Bahkali AH, Kang JC, Wang Y, Wen TC, Wendt L, Xu JC, Camporesi E (2015) Towards unraveling relationships in Xylariomycetidae (Sordariomycetes). Fungal Divers 73:73–144 Senanayake IC, Crous PW, Groenewald JZ, Maharachchikumbura SSN, Jeewon R, Phillips AJL, Bhat JD, Perera RH, Li QR, Li WJ, Tangthirasunun N, Norphanphoun C, Karunarathna SC, Camporesi E, Manawasighe IS, Al-Sadi AM, Hyde KD (2017) Families of Diaporthales based on morphological and phylogenetic evidence. Stud Mycol 86:217–296 Senanayake IC, Jeewon R, Chomnunti P, Wanasinghe DN, Norphanphoun C, Karunarathna A, Pem D, Perera RH, Camporesi E, McKenzie EH, Hyde KD (2018) Taxonomic circumscription of Diaporthales based on multigene phylogeny and morphology. Fungal Divers 93:241–443 Senwanna C, Phookamsak R, Doilom M, Hyde K, Cheewangkoon R (2017) Novel taxa of Diatrypaceae from Para rubber (Hevea brasiliensis) in northern Thailand; introducing anovel genus Allocryptovalsa. Mycosphere 8:1835–1855 Senwanna C, Phookamsak R, Bahkali AH, Elgorban AM, Cheewangkoon R, Hyde KD (2018) Neolinocarpon phayaoense sp. nov. (Linocarpaceae) from Thailand. Phytotaxa 362:77–86 Seto K, Degawa Y (2018) Collimyces mutans gen. et sp. nov. (Rhizophydiales, Collimycetaceae fam. nov.), a new chytrid parasite of Microglena (Volvocales, clade Monadinia). Protist 169:507–520 Seto K, Kagami M, Degawa Y (2017) Phylogenetic position of parasitic chytrids on Diatoms: characterization of a novel clade in Chytridiomycota. J Eukaryot Microbiol 64:383–393 Shang QJ, Hyde KD, Phookamsak R, Doilom M, Bhat DJ, Maharachchikumbura SS, Promputtha I (2017) Diatrypella tectonae and Peroneutypa mackenziei spp. nov. (Diatrypaceae) from northern Thailand. Mycol Prog 16:463–476 123 236 Shearer CA, Raja HA, Miller AN, Nelson P, Tanaka K, Hirayama K, Marvanova L, Hyde KD, Zhang Y (2009) The molecular phylogeny of freshwater Dothideomycetes. Stud Mycol 64:145–153 Shenoy BD, Jeewon R, Wu WP, Bhat DJ, Hyde KD (2006) Ribosomal and RPB2 DNA sequence analyses suggest that Sporidesmium and morphologically similar genera are polyphyletic. Mycol Res 110:916–928 Shoemaker RA (1976) Canadian and some extralimital Ophiobolus species. Can J Bot 54:2365–2404 Shoemaker RA (1984) Canadian and some extralimital Nodulosphaeria and Entodesmium species. Can J Bot 62:2730–2753 Shoemaker RA, Babcock CE (1987) Additions to Nodulosphaeria. Can J Bot 65:1921–1926 Shoemaker RA, Babcock CE (1989) Phaeosphaeria. Can J Bot 67:1500–1599 Shoemaker RA, Müller E (1965) Types of the pyrenomycete genera Hymenopleella and Lepteutypa. Can J Bot 43:1457–1460 Shoemaker R, Hambleton S (2001) ‘‘Helminthosporium’’ asterinum, Polydesmus elegans, Imimyces, and allies. Can J Bot 79:592–599 Sieber TN (2007) Endophytic fungi in forest trees: are they mutualists? Fungal Biol Rev 21:75–89 Silva-Hanlin DMW, Hanlin RT (1999) Small subunit ribosomal RNA gene phylogeny of several loculoascomycetes and its taxonomic implications. Mycol Res 103:153–160 Silvera-Simón C, Mercado-Sierra A, Mena-Portales J, Gene J, Guarro J (2010) Two new species of Digitodesmium from Euskadi (Spain). Cryptogam, Mycol 31:211–218 Singer R, Smith AH (1960) Studies on secotiaceous fungi. IX. The astrogastraceous series. Mem Torrey Bot Club 21:1–122 Sivichai S, Boonyuen N (2010) Jahnula morakotii sp. nov. and J. appendiculata from a peat swamp in Thailand. Mycotaxon 112:475–481 Slippers B, Wingfield MJ (2007) Botryosphaeriaceae as endophytes and latent pathogens of woody plants: diversity, ecology and impact. Fungal Biol Rev 21:90–106 Slippers B, Boissin E, Phillips AJL, Groenewald JZ, Lombard L, Wingfield MJ, Postma A, Burgess T, Crous PW (2013) Phylogenetic lineages in the Botryosphaeriales: a systematic and evolutionary framework. Stud Mycol 76:31–49 Slippers B, Roux J, Wingfield MJ, van der Walt FJJ, Jami F, Mehl JWM, Marais GJ (2014) Confronting the constraints of morphological taxonomy in the Botryosphaeriales. Persoonia 33:155–168 Smith SE, Read D (2008) Mycorrhizal symbiosis, 3rd edn. Academic Press, Boston Soares M, Kritzberg ES, Rousk J (2017) Labile carbon ‘primes’ fungal use of nitrogen from submerged leaf litter. FEMS Microbiol Ecol. https://doi.org/10.1093/femsec/fix110 Sobreira PNB, Cáceres MES, Maia LC, Lücking R (2018) Flabelloporina, a new genus in the Porinaceae (Ascomycota, Ostropales), with the first record of F. squamulifera from Brazil. Phytotaxa 358:67–75 Sogonov MV, Castlebury LA, Rossman AY, Mejı́a LC, White JF (2008) Leaf-inhabiting genera of the Gnomoniaceae, Diaporthales. Stud Mycol 62:1–77 Spatafora JW, Mitchell TG, Vilgalys R (1995) Analysis of genes coding for small-subunit rRNA sequences in studying phylogenetics of dematiaceous fungal pathogens. J Clin Microbiol 33:1322–1326 Spegazzini CL (1911) Mycetes Argentinenses. Series V. Anales Mus Nac Hist Nat Buenos Aires 3:446 Sri-indrasutdhi V, Boonyuen N, Suetrong S, Chuaseeharonnachai C, Sivichai S, Jones EBG (2010) Wood-inhabiting freshwater fungi from Thailand: Ascothailandia grenadoidia gen. et sp. nov., Canalisporium grenadoidia sp. nov. with a key to 123 Fungal Diversity (2019) 96:1–242 Canalisporium species (Sordariomycetes, Ascomycota). Mycoscience 51:411–420 Stamatakis A (2014) RAxML Version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30:1312–1313 Steyaert RL (1949) Contribution à l’étude monographigue de Pestalotia de Not. et Monochaetia Sacc. (Truncatella gen. nov. et Pestalotiopsis gen. nov.). Bull Jard Bot Brux 19:285–354 Studer A, Nusbaumer L, Spichiger R (2015) Biodiversidade da Reserva Biológica de Pedra Talhada Alagoas, PernambucoBrasil. Boissiera 68:1–818 Sturm J (1829) Deutschlands Flora. Abt III Die Pilze Deutschl 2:1–136 Su H, Hyde KD, Maharachchikumbura SSN, Ariyawansa HA, Luo Z, Promputtha I, Tian Q, Lin C, Shang Q, Zhao Y, Chai H, Liu X, Bahkali AH, Bhat JD, McKenzie EHC, Zhou D (2016) The families Distoseptisporaceae fam. nov., Kirschsteiniotheliaceae, Sporormiaceae and Torulaceae, with new species from freshwater in Yunnan Province, China. Fungal Divers 80:375–409 Subramanian CV (1992) A reassessment of Sporidesmium (Hyphomycetes) and some related taxa. Proc Indian Acad Sci (Plant Sciences) 58(4):179–190 Subramanian CV (1994) Hyphomycetes from South East Asia— novelties from Singapore and Malaysia. Kavaka 22:52–76 Suetrong S, Schoch CL, Spatafora JW, Kohlmeyer J, VolkmannKohlmeyer B, Sakayaroj J, Phongpaichit S, Tanaka K, Hirayama Jones EBG (2009) Molecular systematics of the marine Dothideomycetes. Stud Mycol 64:155–173 Suetrong S, Boonyuen N, Pang KL, Ueapattanakit J, Klaysuban A, Sriiindrasutdhi V, Sivichai S, Jones EBG (2011a) A taxonomic revision and phylogenetic reconstruction of the Jahnulales (Dothideomycetes), and the new family Manglicolaceae. Fungal Divers 51:163–188 Suetrong S, Hyde KD, Ahang Y, Bahkali AB, Jones EBG (2011b) Trematosphaeriaceae fam. nov. (Dothideomycetes, Ascomycota). Cryptogam, Mycol 32:343–358 Suetrong S, Rungjindamai N, Sommai S, Rung-Areerate P, Sommrithipol S, Jones EBG (2014) Wiesneriomyces a new lineage of Dothideomycetes (Ascomycota) basal to Tubeufiales. Phytotaxa 176:283–297 Summerbell RC, Gueidan C, Schroers HJ, de Hoog GS, Starink M, Arocha Rosete Y, Guarro J, Scott JA (2011) Acremonium phylogenetic overview and revision of Gliomastix, Trichothecium and Sarocladium. Stud Mycol 68:139–162 Sung GH, Spatafora JW (2004) Cordyceps cardinalis sp. nov., a new species of Cordyceps with an east Asian-eastern North American distribution. Mycologia 96:658–666 Sung GH, Hywel-Jones NL, Sung JM, Luangsa-ard JJ, Shrestha B, Spatafora JW (2007) Phylogenetic classification of Cordyceps and the clavicipitaceous fungi. Stud Mycol 57:5–59 Sutton BC (1973) Coelomycetes VI. Nomenclature of generic names proposed for Coelomycetes. Mycol Pap 141:1–253 Sutton BC (1980) The Coelomycetes. Fungi imperfecti with pycnidia, acervuli and stromata. Ann Mycol 6:484 Svrcek M (1977) New combinations and new taxa in operculate Discomycetes (Pezizales). Ceská Mykol 31:69–71 Tanaka K, Harada Y (2003) Pleosporales in Japan (1): the genus Lophiostoma. Mycoscience 44:85–96 Tanaka K, Hirayama K, Yonezawa H, Hatakeyama S, Harada Y, Sano T, Shirouzu T, Hosoya T (2009) Molecular taxonomy of bambusicolous fungi: Tetraplosphaeriaceae, a new pleosporalean family with Tetraploa-like anamorphs, and notes on the phylogeny of selected species from bamboo. Stud Mycol 64:175–209 Tanaka K, Hirayama K, Yonezawa H, Sato G, Toriyabe A, Kudo H, Hashimoto A, Matsumura M, Harada Y, Kurihara Y, Shirouzu T Fungal Diversity (2019) 96:1–242 (2015) Revision of the Massarineae (Pleosporales, Dothideomycetes). Stud Mycol 82:75–136 Tang AMC, Jeewon R, Hyde KD (2009) A re-evaluation of the evolutionary relationships within the Xylariaceae based on ribosomal and protein-coding gene sequences. Fungal Divers 34:127–155 Tanney J, Miller AN (2017) Asexual-sexual morph connection in the type species of Berkleasmium. IMA Fungus 8:99–105 Tavares II (1985) Laboulbeniales (Fungi, Ascomycetes). Mycol Mem 9:1–627 Tedersoo L, Nara K (2010) General latitudinal gradient of biodiversity is reversed in ectomycorrhizal fungi. New Phytol 185:351–354 Tedersoo L, Sánchez-Ramı́rez S, Koljalg U, Bahram M, Döring M, Schigel D, May T, Ryberg M, Abarenkov K (2018) High-level classification of the fungi and a tool for evolutionary ecological analyses. Fungal Divers 90:135–159 Teixeira MdM, Moreno LF, Stielow BJ, Muszewska A, Hainaut M, Gonzaga L, Abouelleil A, Patané JSL, Priest M, Souza R, Young S, Ferreira KS, Zeng Q, da Cunha MML, Gladki A, Barker B, Vicente VA, de Souza EM, Almeida S, Henrissat B, Vasconcelos ATR, Deng S, Voglmayr H, Moussa TAA, Gorbushina A, Felipe MSS, Cuomo CA, de Hoog GS (2017) Exploring the genomic diversity of black yeasts and relatives (Chaetothyriales, Ascomycota). Stud Mycol 86:1–28 Teng SC (1936) Additional fungi from China IV. Sinensia 7:752–823 Tennakoon DS, Hyde KD, Phookamsak R, Wanasinghe DN, Camporesi E, Promputtha I (2016a) Taxonomy and Phylogeny of Juncaceicola gen. nov. (Phaeosphaeriaceae, Pleosporinae, Pleosporales). Cryptogam, Mycol 37:135–156 Tennakoon DS, Phillips AJL, Phookamsak R, Ariyawansa HA, Bahkali AH, Hyde KD (2016b) Sexual morph of Lasiodiplodia pseudotheobromae (Botryosphaeriaceae, Botryosphaeriales, Dothideomycetes) from China. Mycosphere 7:990–1000 Tennakoon DS, Kuo CH, Jeewon R, Thambugala KM, Hyde KD (2018) Saprobic Lophiostomataceae (Dothideomycetes): Pseudolophiostoma mangiferae sp. nov. and Neovaginatispora fuckelii, a new record from Mangifera indica. Phytotaxa 364:157–171 Thambugala KM, Hyde KD (2018) Additions to the genus Massariothea in Diaporthaceae. Mycol Prog 17:1139–1147 Thambugala KM, Hyde KD, Tanaka K, Tian Q, Wanasinghe DN, Ariyawansa HA, Jayasiri SC, Boonmee S, Camporesi E, Hashimoto A, Hirayama K (2015) Towards a natural classification and backbone tree for Lophiostomataceae, Floricolaceae and Amorosiaceae fam. nov. Fungal Divers 74:199–266 Thambugala KM, Daranagama DA, Phillips AJL, Bulgakov TS, Bhat DJ, Camporesi E, Bahkali AH, Eungwanichayapant PD, Liu ZY, Hyde KD (2017a) Microfungi on Tamarix. Fungal Divers 82:239–306 Thambugala KM, Wanasinghe DN, Phillips AJL, Camporesi E, Bulgakov TS, Phukhamsakda C, Ariyawansa HA, Goonasekara ID, Phookamsak R, Dissanayake A, Tennakoon DS, Tibpromma S, Chen YY, Liu ZY, Hyde KD (2017b) Mycosphere notes 1-50: Grass (Poaceae) inhabiting Dothideomycetes. Mycosphere 8:697–796 Thambugala KM, Hyde KD, Zhang JF, Liu ZY (2018) Didymella eriobotryae sp. nov. (Didymellaceae) and Arthrinium arundinis (Apiosporaceae) from fruit of Eriobotrya japonica (loquat) in China. Phytotaxa 382:136–147 Thaxter R (1908) Contribution toward a monograph of the Laboulbeniaceae. Part II. Mem Am Acad Arts Sci 13:217–469 Thaxter R (1917) New Laboulbeniales, chiefly dipterophilous American species. Proc Am Acad Arts Sci 52:647–721 Thaxter R (1918) Extra American Dipterophilous Laboulbeniales. Proc Am Acad Arts Sci 53:697–749 237 Thaxter R (1931) Contribution towards a monograph of the Laboulbeniaceae. Part V. Mem Am Acad Arts Sci 16:1–435 Theissen F (1913) Lembosia – Studien. Ann Mycol 11:424–467 Tian Q, Doilom M, Luo ZL, Chomnunti P, Bhat JD, Xu JC, Hyde KD (2016) Introducing Melanoctona tectonae gen. et sp. nov. and Minimelanolocus yunnanensis sp. nov. (Herpotrichiellaceae, Chaetothyriales). Cryptogam, Mycol 37:477–492 Tibpromma S, Bhat J, Doilom M, Lumyong S, Nontachaiyapoom S, Yang JB, Hyde KD (2016) Three new Hermatomyces species (Lophiotremataceae) on Pandanus odorifer from Southern Thailand. Phytotaxa 275:127–139 Tibpromma S, Hyde KD, Jeewon R, Maharachchikumbura SS, Liu JK, Bhat DJ, Jones EBG, McKenzie E, Camporesi E, Bulgakov TS, Doilom M, Santiago AM, Das K, Manimohan P, Gibertoni TB, Lim YW, Ekanayaka AH, Thongbai B, Lee HB, Yang J, Kirk PM, Sysouphanthong P, Singh SK, Boonmee S, Dong W, Raj KN, Latha KP, Phookamsak R, Phukhamsakda C, Konta S, Jayasiri SC, Norphanphoun C, Tennakoon D, Li J, Dayarathne MC, Perera RH, Xiao Y, Wanasinghe DN, Senanayake IC, Goonasekara ID, Silva NI, Mapook A, Jayawardena RS, Dissanayake AJ, Manawasinghe IS, Chethana KW, Luo Z, Hapuarachchi KK, Baghela A, Soares AM, Vizzini A, MeirasOttoni A, Mešić A, Dutta AK, Souza CA, Richter C, Lin C, Chakrabarty D, Daranagama DA, Lima DX, Chakraborty D, Ercole E, Wu F, Simonini G, Vasquez G, Silva GA, Plautz HL, Ariyawansa HA, Lee HS, Kušan I, Song J, Sun J, Karmakar J, Hu K, Semwal KC, Thambugala KM, Voigt K, Acharya K, Rajeshkumar KC, Ryvarden L, Jadan M, Hosen MI, Mikšı́k M, Samarakoon MA, Wijayawardene NN, Kim NK, Matočec N, Singh PN, Tian Q, Bhatt RP, Oliveira RJ, Tulloss RE, Aamir S, Kaewchai S, Marathe SD, Khan S, Hongsanan S, Adhikari S, Mehmood T, Bandyopadhyay TK, Svetasheva TY, Nguyen TT, Antonı́n V, Li W, Wang Y, Indoliya Y, Tkalčec Z, Elgorban AM, Bahkali AH, Tang A, Su H, Zhang H, Promputtha I, Luangsa-Ard J, Xu J, Yan J, Kang JC, Stadler M, Mortimer PE, Chomnunti P, Zhao Q, Phillips AJ, Nontachaiyapoom S, Wen T, Karunarathna SC (2017) Fungal diversity notes 491–602: taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 83:1–261 Tibpromma S, Hyde KD, McKenzie EH, Bhat DJ, Phillips AJ, Wanasinghe DN, Samarakoon MC, Jayawardena RS, Dissanayake AJ, Tennakoon DS, Doilom M, Phookamsak R, Tang AMC, Xu J, Mortimer PE, Promputtha I, Maharachchikumbura SSN, Khan S, Karunarathna SC (2018) Fungal diversity notes 840–928: micro-fungi associated with Pandanaceae. Fungal Divers 93:1–60 Timdal E (1991) A monograph of the genus Toninia (Lecideaceae, Ascomycetes). Opera Bot 110:1–137 Trouillas FP, Pitt WM, Sosnowski MR, Huang R, Peduto F, Loschiavo A, Savocchia S, Scott ES, Gubler WD (2011) Taxonomy and DNA phylogeny of Diatrypaceae associated with Vitis vinifera and other woody plants in Australia. Fungal Divers 49:203–223 Tubaki K (1958) Studies on the Japanese hyphomycetes. V. Leaf & stem group with a discussion of the classification of the hyphomycetes and their perfect stages. J Hattori Bot Lab 20:142–244 Udayanga D, Castlebury LA, Rossman AY, Hyde KD (2014) Species limits in Diaporthe: molecular re-assessment of D. citri, D. cytosporella, D. foeniculina and D. rudis. Persoonia 32:83–101 Úrbez-Torres JR, Boule J, O’Gorman DT (2016) First report of Diplodia seriata and D. mutila causing apple dieback in British Columbia. Plant Dis 100:1243–1244 Valdez JG, Makuch MA, Ordovini AF, Masuelli RW, Overy DP, Piccolo RJ (2006) First report of Penicillium allii as a field pathogen of garlic (Allium sativum). Plant Pathol 55:583 123 238 Valenzuela-Lopez N, Cano-Lira JF, Guarro J, Sutton DA, Wiederhold N, Crous PW, Stchigel AM (2018) Coelomycetous Dothideomycetes with emphasis on the families Cucurbitariaceae and Didymellaceae. Stud Mycol 90:1–69 Van den Wyngaert S, Seto K, Rojas-Jimenez K, Kagami M, Grossart HP (2017) A new parasitic chytrid, Staurastromyces oculus (Rhizophydiales, Staurastromycetaceae fam. nov.), infecting the freshwater desmid Staurastrum sp. Protist 168:392–407 Vasilyeva LN, Stephenson SL (2005) Pyrenomycetes of the Great Smoky Mountains National Park. II. Cryptovalsa Ces. et De Not. and Diatrypella (Ces. et De Not.) Nitschke (Diatrypaceae). Fungal Divers 19:189–200 Verbeken A, Nuytinck J (2013) Not every milkcap is a Lactarius. Scr Bot Belg 51:162–168 Verma RK, Sharma N, Soni KK, Jamaluddin (2008) Forest fungi of Central India. International Book Distributing Co., Lucknow, pp 11–418 Vesterholt J (2012) Coprinopsis P. Karst. In: Knudsen H, Vesterholt J (eds) Funga Nordica. Nordsvamp, Copenhagen, pp 672–687 Visagie CM, Houbraken J, Frisvad JC, Hong SB, Klaassen CHW, Perrone G, Seifert KA, Varga J, Yaguchi T, Samson RA (2014) Identification and nomenclature of the genus Penicillium. Stud Mycol 78:343–371 Voglmayr H, Jaklitsch WM (2011) Molecular data reveal high host specificity in the phylogenetically isolated genus Massaria (Ascomycota, Massariaceae). Fungal Divers 46:133–170 Voglmayr H, Friebes G, Gardiennet A, Jaklitsch WM (2018) Barrmaelia and Entosordaria in Barrmaeliaceae (fam. nov., Xylariales) and critical notes on anthostomella-like genera based on multigene phylogenies. Mycol Prog 17:155–177 Von Arx JA, Müller E (1975) A re-evaluation of the bitunicate ascomycetes with keys to families and genera. Stud Mycol 9:1–159 von Höhnel F (1909) Fragmente zur Mykologie: VI. Mitteilung (Nr. 182 bis 288). Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften Math.-naturw. Klasse Abt I 118:275–452 Vu D, Groenewald M, de Vries M, Gehrmann T, Stielow B, Eberhardt U, Al-Hatmi A, Groenewald JZ, Cardinali G, Houbraken J, Boekhout T, Crous PW, Robert V, Verkley GJM (2019) Largescale generation and analysis of filamentous fungal DNA barcodes boosts coverage for kingdom fungi and reveals thresholds for fungal species and higher taxon delimitation. Stud Mycol 92:135–154 Walker DM, Lawrence BR, Wooten JA, Rossman AY, Castlebury LA (2014) Five new species of the highly diverse genus Plagiostoma (Gnomoniaceae, Diaporthales) from Japan. Mycol Prog 13:1057–1067 Wanasinghe DN, Jones EBG, Camporesi E, Mortimer PE, Xu JC, Bahkali AH, Hyde KD (2015) The genus Murispora. Cryptogam, Mycol 36:419–448 Wanasinghe DN, Jones EBG, Dissanayake AJ, Hyde KD (2016) Saprobic Dothideomycetes in Thailand: Vaginatispora appendiculata sp. nov. (Lophiostomataceae) introduced based on morphological and molecular data. Stud Fungi 1:56–58 Wanasinghe DN, Hyde KD, Jeewon R, Crous PW, Wijayawardene NN, Jones EBG, Phukhamsakda C (2017a) Phylogenetic revision of Camarosporium (Pleosporineae, Dothideomycetes) and allied genera. Stud Mycol 87:207–256 Wanasinghe DN, Phookamsak R, Jeewon R, Li Wen Jing, Hyde KD, Jones EBG, Camporesi E, Promputtha I (2017b) A family level rDNA based phylogeny of Cucurbitariaceae and Fenestellaceae with descriptions of new Fenestella species and Neocucurbitaria gen. nov. Mycosphere 8:397–414 Wanasinghe DN, Phukhamsakda C, Hyde KD, Jeewon R, Lee HB, Jones EBG, Tibpromma S, Tennakoon DS, Dissanayake AJ, Jayasiri SC, Gafforov Y, Camporesi E, Bulgakov TS, Ekanayake 123 Fungal Diversity (2019) 96:1–242 AH, Perera RH, Samarakoon MC, Goonasekara ID, Mapook A, Li WJ, Senanayake IC, Li JF, Norphanphoun C, Doilom M, Bahkali AH, Xu JC, Mortimer PE, Tibell L, Tibell S, Karunarathna SC (2018a) Fungal diversity notes 709–839: taxonomic and phylogenetic contributions to fungal taxa with an emphasis on fungi on Rosaceae. Fungal Divers 89:1–236 Wanasinghe DN, Jeewon R, Peršoh D, Jones EBG, Camporesi E, Bulgakov TS, Gafforov YS, Hyde KD (2018b) Taxonomic circumscription and phylogenetics of novel didymellaceous taxa with brown muriform spores. Stud Fungi 3:152–175 Wang RX, Luo ZL, Hyde KD, Bhat DJ, Su XJ, Su HY (2016) New species and records of Dictyocheirospora from submerged wood in north-western Yunnan, China. Mycosphere 7:357–1367 Webster J (2007) Introduction to fungi. VIII ? 424 S., 242 Abb. Cambridge University Press, Cambridge Wehmeyer LE (1926) A biologic and phylogenetic study of the stromatic Sphaeriales. Am J Bot 13:575–645 Wei MJ, Zhang H, Dong W, Boonmee S, Zhang D (2018) Introducing Dictyochaeta aquatica sp. nov. and two new species of Chloridium (Chaetosphaeriaceae, Sordariomycetes) from aquatic habitats. Phytotaxa 362:187–199 Wendt L, Sir EB, Kuhnert E, Heitkämper S, Lambert C, Hladki AI, Romero AI, Luangsa-ard JJ, Srikitikulchai P, Peršoh D, Stadler M (2018) Resurrection and emendation of the Hypoxylaceae, recognised from a multigene phylogeny of the Xylariales. Mycol Prog 17:115–154 Weresub LK, LeClair PM (1971) On Papulaspora and bulbilliferous basidiomycetes Burgoa and Minimedusa. Can J Bot 49:2203–2213 Wijayawardene NN, Crous PW, Kirk PM, Hawksworth DL, Boonmee S, Braun U, Dai DQ, D’souza MJ, Diederich P, Dissanayake A, Doilom M (2014) Naming and outline of Dothideomycetes–2014 including proposals for the protection or suppression of generic names. Fungal Divers 69:1–55 Wijayawardene NN, Hyde KD, Rajeshkumar KC, Hawksworth DL, Madrid H, Kirk PM, Braun U, Singh RV, Crous PW, Kukwa M, Lucking R, Kurtzman CP, Yurkov A, Haelewaters D, Aptroot A, Lumbsch HT, Timdal E, Ertz D, Etayo J, Phillips AJL, Groenewald JZ, Papizadeh M, Selbmann L, Dayarathne MC, Weerakoon G, Jones EBG, Suetrong S, Tian Q, Castanéda-Ruiz RF, Bahkali AH, Pang KL, Tanaka K, Dai DQ, Sakayaroj J, Hujslová M, Lombard L, Shenoy BD, Suija A, Maharachchikumbura SSN, Thambugala KM, Wanasinghe DN, Sharma BO, Gaikwad S, Pandit G, Zucconi L, Onofri S, Egidi E, Raja HA, Kodsueb R, Caceres MES, Perez-Ortega S, Fiuza PO, Monteiro JS, Vasilyeva LN, Shivas RG, Prieto M, Wedin M, Olariaga I, Lateef AA, Agrawal Y, Fazeli SAS, Amoozegar MA, Zhao GZ, Pfliegler WP, Sharma G, Oset M, Abdel MA, Takamatsu S, Bensch K, Silva NI, De Kesel A, Karunarathna A, Boonmee S, Pfister DH, Lu YZ, Luo ZL, Boonyuen N, Daranagama DA, Senanayake IC, Jayasiri SC, Samarakoon MC, Zeng XY, Doilom M, Quijada L, Rampadarath S, Heredia G, Dissanayake AJ, Jayawardana RS, Perera PH, Tang LZ, Phukhamsakda C, Hernández-Restrepo M, Ma XY, Tibpromma S, Gusmao LFP, Weerahewa D, Karunarathna SC (2017a) Notes for genera: Ascomycota. Fungal Divers 86:1–594 Wijayawardene NN, Hyde KD, Tibpromma S, Wanasinghe DN, Thambugala KM, Tian Q, Wang Y (2017b) Towards incorporating asexual fungi in a natural classification: checklist and notes 2012–2016. Mycosphere 8:1457–1554 Wijayawardene NN, Hyde KD, Lumbsch HT, Liu JK, Maharachchikumbura SSN, Ekanayaka AH, Tian Q, Phookamsak R (2018a) Outline of Ascomycota: 2017. Fungal Divers 88:167–263 Wijayawardene NN, Pawłowska J, Letcher PM, Kirk PM, Humber RA, Schüßler A, Wrzosek M, Muszewska A, Okrasińska A, Istel Fungal Diversity (2019) 96:1–242 Ł, Ge˛siorska A, Mungai P, Azeez Lateef A, Rajeshkumar KC, Singh RV, Radek R, Walther G, Wagner L, Walker C, Wijesundara DSA, Papizadeh M, Dolatabadi S, Shenoy BD, Tokarev YS, Lumyong S, Hyde KD (2018b) Notes for genera: basal clades of Fungi (including Aphelidiomycota, Basidiobolomycota, Blastocladiomycota, Calcarisporiellomycota, Caulochytriomycota, Chytridiomycota, Entomophthoromycota, Glomeromycota, Kickxellomycota, Monoblepharomycota, Mortierellomycota, Mucoromycota, Neocallimastigomycota, Olpidiomycota, Rozellomycota and Zoopagomycota. Fungal Divers 92:43–129 Winter HG (1885) Pilze. II. Abtheilung. Ascomyceten: Gymnoasceen und Pyrenomyceten. In: Rabenhorst GL (ed) Rabenhorst’s Kryptogamen-Flora von Deutschland, Oesterreich und der Schweiz, 1(2). Kummer, Leipzig, pp 65–528 Wu F, Zhou L, Ji X, Tian X, He S (2016) Grammothele hainanensis sp. nov. (Polyporales, Basidiomycota) and related species from Hainan, southern China. Phytotaxa 255:160–166 Xia JW, Ma YR, Li Z, Zhang XG (2017) Acrodictys-like wood decay fungi from southern China, with two new families Acrodictyaceae and Junewangiaceae. Sci Rep 7:7888 Yang J, Liu JK, Hyde KD, Bhat DJ, Jones EBG, Liu ZY (2016a) New species of Sporoschisma (Chaetosphaeriaceae) from aquatic habitats in Thailand. Phytotaxa 289:147–157 Yang T, Groenewald JZ, Cheewangkoon R, Jami F, Abdollahzadeh J, Lombard L, Crous PW (2016b) Families, genera and species of Botryosphaeriales. Fungal Biol 121:322–346 Yang P, Shi W, Wang H, Liu H (2016c) Screening of freshwater fungi for decolorizing multiple synthetic dyes. Braz J Microbiol 47:828–834 Yang J, Liu NG, Liu JK, Hyde KD, Jones EBG, Liu ZY (2018a) Phylogenetic placement of Cryptophiale, Cryptophialoidea, Nawawia, Neonawawia gen. nov. and Phialosporostilbe. Mycosphere 9:1132–1150 Yang J, Maharachchikumbura SSN, Liu JK, Hyde KD, Jones EBG, Al-Sadi AM, Liu ZY (2018b) Pseudostanjehughesia aquitropica gen. et sp. nov. and Sporidesmium sensu lato species from freshwater habitats. Mycol Prog 17:591–616 Yao YJ, Spooner BM (1996) Delimitation of Boubovia and Pulvinula. Mycol Res 100:193–194 Yeh YH, Kirschner R (2014) Sarocladium spinificis, a New Endophyte Species from the Coastal Grass Spinifex littoreus in Taiwan. Bot Stud 55:25 Ying JZ (1983) A study on Russula virdi-rubrolimbata sp. nov. and its related species of subsection Virescentinas. Mycosystema 2:34–37 Yuan Z, Zhao Z (1994) Studies on lophiostomataceous fungi from Xinjiang, China. Sydowia 46:162–184 Yun YH, Oh MH (2016) First report of Nothophoma quercina causing shoot canker on Chaenomeles sinensis in Korea. Plant Dis 100:2533 Zang M, Kinjo N (1998) Notes on the alpine Cordyceps of China and nearby nations. Mycotaxon 66:215–230 Zeng XY, Jeewon R, Wen TC, Hongsanan S, Boonmee S, Hyde KD (2018) Simplified and efficient DNA extraction protocol for Meliolaceae specimens. Mycol Prog 17:403–415 Zhang N, Wang Z (2015) Pezizomycotina: Sordariomycetes and Leotiomycetes. In: McLaughlin DJ, Spatafora JW (eds) The Mycota. Part B, Systematics and evolution. Springer, Berlin, pp 57–88 239 Zhang N, Castlebury LA, Miller AN, Huhndorf SM, Schoch CL, Seifert KA, Rossman AY, Rogers JD, Kohlmeyer J, VolkmannKohlmeyer B, Sung GH (2006) An overview of the systematics of the Sordariomycetes based on a four-gene phylogeny. Mycologia 98:1076–1087 Zhang Y, Jeewon R, Fournier J, Hyde KD (2008) Multi-gene phylogeny and morphotaxonomy of Amniculicola lignicola: a novel freshwater fungus from France and its relationships to the Pleosporales. Mycol Res 112:1186–1194 Zhang Y, Schoch CL, Fournier J, Crous PW, de Gruyter J, Woudenberg JHC, Hirayama K, Tanaka K, Pointing SB, Spatafora JW, Hyde KD (2009a) Multi–locus phylogeny of Pleosporales: a taxonomic, ecological and evolutionary re– evaluation. Stud Mycol 64:85–102 Zhang Y, Fournier J, Crous PW, Pointing SB, Hyde KD (2009b) Phylogenetic and morphological assessment of two new species of Amniculicola and their allies (Pleosporales). Persoonia 23:48–54 Zhang Y, Wang HK, Fournier J, Crous PW (2009c) Towards a phylogenetic clarification of Lophiostoma/Massarina and morphologically similar genera in the Pleosporales. Fungal Divers 38:225–251 Zhang Y, Crous PW, Schoch CL, Hyde KD (2012) Pleosporales. Fungal Divers 53:1–221 Zhang H, Hyde KD, Zhao YC, McKenzie EHC (2014a) Freshwater ascomycetes: Lophiostoma vaginatispora comb. nov. (Dothideomycetes, Pleosporales, Lophiostomaceae) based on morphological and molecular data. Phytotaxa 176:184–191 Zhang Q, Zhang J, Yang L, Zhang L, Jiang D, Chen W, Li G (2014b) Diversity and biocontrol potential of endophytic fungi in Brassica napus. Biol Control 72:98–108 Zhang JB, Li JW, Li F, Qiu LH (2017) Russula dinghuensis sp. nov. and R. subpallidirosea sp. nov., two new species from southern China supported by morphological and molecular evidence. Cryptogam, Mycol 38:191–203 Zhao Q, Li YK, Zhu XT, Zhao YC, Liang JF (2015) Russula nigrovirens sp. nov. (Russulaceae) from southwestern China. Phytotaxa 236:249–256 Zhao RL, Li GJ, Sánchez-Ramı́rez S, Stata M (2017) A six-gene phylogenetic overview of Basidiomycota and allied phyla with estimated divergence times of higher taxa and a phyloproteomics perspective. Fungal Divers 84:43–74 Zhou LW, Vlasák J, Dai YC (2016a) Taxonomy and phylogeny of Phellinidium (Hymenochaetales, Basidiomycota): a redefinition and the segregation of Coniferiporia gen. nov. for forest pathogens. Fungal Biol 120:988–1001 Zhou LW, Vlasák J, Decock C, Assefa A, Stenlid J, Abate D, Wu SH, Dai YC (2016b) Global diversity and taxonomy of the Inonotus linteus complex (Hymenochaetales, Basidiomycota): Sanghuangporus gen. nov., Tropicoporus excentrodendri and T. guanacastensis gen. et spp. nov., and 17 new combinations. Fungal Divers 77:335–347 Zhou LW, Vlasák J, Qin WM, Dai YC (2016c) Global diversity and phylogeny of the Phellinus igniarius complex (Hymenochaetales, Basidiomycota) with the description of five new species. Mycologia 108:192–204 Zlatković M, Keča N, Wingfield MJ, Jami F, Slippers B (2016) New and unexpected host associations for Diplodia sapinea in the Western Balkans. For Pathol 47(3):1–11 123 240 Fungal Diversity (2019) 96:1–242 Affiliations Kevin D. Hyde1,2,3,4,5,6,71 • Danushka S. Tennakoon1,2,3,6,7 • Rajesh Jeewon8 • D. Jayarama Bhat9,10 • Sajeewa S. N. Maharachchikumbura11 • Walter Rossi12 • Marco Leonardi12 • Hyang Burm Lee13 • Hye Yeon Mun14 • Jos Houbraken15 • Thuong T. T. Nguyen13 • Sun Jeong Jeon13 • Jens Christian Frisvad16 • Dhanushka N. Wanasinghe1,3,4,71 • Robert Lücking17 • André Aptroot18 • Marcela E. S. Cáceres19 • Samantha C. Karunarathna1,4,5,71 • Sinang Hongsanan3,20 • Rungtiwa Phookamsak1,3,4,5,71 • Nimali I. de Silva1,3,5 • Kasun M. Thambugala21 • Ruvishika S. Jayawardena3 • Indunil C. Senanayake3,20 • Saranyaphat Boonmee3 • Jie Chen22 • Zong-Long Luo23 • Chayanard Phukhamsakda2,3 • Olinto L. Pereira24 • Vanessa P. Abreu25 • André Wilson Campos Rosado24 • Buyck Bart26 • Emile Randrianjohany27 • Valérie Hofstetter28 • Tatiana B. Gibertoni29 • Adriene Mayra da Silva Soares30 • Helio Longoni Plautz Jr.31 • Helen Maria Pontes Sotão30 • William Kalhy Silva Xavier32 • Jadson Diogo Pereira Bezerra33 • Thays Gabrielle Lins de Oliveira33 • Cristina Maria de Souza-Motta33 • Oliane Maria Correia Magalhães33 • Digvijayini Bundhun3,34 • Dulanjalee Harishchandra2,3,35 • Ishara S. Manawasinghe2,3,35 • Wei Dong3,6,34,36 • Sheng-Nan Zhang3,34 • Dan-Feng Bao3,23,34 • Milan C. Samarakoon3,5,37 • Dhandevi Pem2,3,6,20 • Anuruddha Karunarathna1,3,7,34 • Chuan-Gen Lin2,3,6 • Jing Yang2,3,6,37 • Rekhani H. Perera2,3,6,37 • Vinit Kumar3,34 • Shi-Ke Huang1,2,3,6 • Monika C. Dayarathne1,2,3,6 • Anusha H. Ekanayaka1,2,3 • Subashini C. Jayasiri1,3 • Yuanpin Xiao2,3,6,38 • Sirinapa Konta1,2,3,6 • Tuula Niskanen39 • Kare Liimatainen39 • Yu-Cheng Dai40 • Xiao-Hong Ji40 • Xue-Mei Tian41 • Armin Mešić42 • Sanjay K. Singh43 • Kunthida Phutthacharoen2,3,6 • Lei Cai4 • Touny Sorvongxay3 • Vinodhini Thiyagaraja1,3,6,34 • Chada Norphanphoun2,3,6,7,38 • Napalai Chaiwan1,2,3,6 • Yong-Zhong Lu3,6,38 • Hong-Bo Jiang1,2,3,6 • Jin-Feng Zhang3,37 • Pranami D. Abeywickrama2,3,35 • Janith V. S. Aluthmuhandiram2,3,35 • Rashika S. Brahmanage2,3,35 • Ming Zeng1,2,3,6 • Thilini Chethana2,3,35 • Deping Wei1,3,34 • Martina Réblová45 • Jacques Fournier46 • Jana Nekvindová47 • Renan do Nascimento Barbosa48 • José Ewerton Felinto dos Santos33 • Neiva Tinti de Oliveira33 • Guo-Jie Li44 • Damien Ertz49,50 • Qiu-Ju Shang2,3,37 • Alan J. L. Phillips51 • Chang-Hsin Kuo7 • Erio Camporesi52,53,54 • Timur S. Bulgakov55 • Saisamorn Lumyong3,5,68,69 • E. B. Gareth Jones3,56 • Putarak Chomnunti2,3 • Eleni Gentekaki2,3 • Frank Bungartz57,58,59 • Xiang-Yu Zeng3,38 • Sally Fryar60 • Zdenko Tkalčec42 • Junmin Liang44 • Guangshuo Li44,61 • Ting-Chi Wen38,62 • Paras Nath Singh43 • Yusufjon Gafforov63,64,70 • Itthayakorn Promputtha5,72 • Erandi Yasanthika2,3 • Ishani D. Goonasekara1,2,3 • Rui-Lin Zhao44 • Qi Zhao1 • Paul M. Kirk65 • Jian-Kui Liu37,66 • JiYe Yan35 • Peter E. Mortimer1,71 • Jianchu Xu1,4,71 • Mingkwan Doilom1,3,4,5,67,71 1 Key Laboratory for Plant Diversity and Biogeography of East Asia, Kunming Institute of Botany, Chinese Academy of Science, Kunming 650201, Yunnan, People’s Republic of China 10 No. 128/1-J, Azad Housing Society, Curca, Goa Velha 403108, India 11 Department of Crop Sciences, College of Agricultural and Marine Sciences, Sultan Qaboos University, PO Box 8, 123 Al Khoud, Oman 2 School of Science, Mae Fah Luang University, Chiang Rai 57100, Thailand 12 3 Center of Excellence in Fungal Research, Mae Fah Luang University, Chiang Rai 57100, Thailand Sect. Environmental Sciences, Dept. MeSVA, University of L’Aquila, 67100 Coppito, AQ, Italy 13 4 East and Central Asia Regional Office, World Agroforestry Centre (ICRAF), Kunming 650201, Yunnan, People’s Republic of China Division of Food Technology, Biotechnology & Agrochemistry, College of Agriculture & Life Sciences, Chonnam National University, Gwangju 61186, Korea 14 5 Department of Biology, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand Fungal Resources Research Division, NNIBR, Sangju-si 37242, Korea 15 6 Mushroom Research Foundation, 128 M.3 Ban Pa Deng T. Pa Pae, A. Mae Taeng, Chiang Mai 50150, Thailand Westerdijk Fungal Biodiversity Institute, Uppsalalaan 8, 3584 CT Utrecht, The Netherlands 16 7 Department of Plant Medicine, National Chiayi University, 300 Syuefu Road, Chiayi City 60004, Taiwan, People’s Republic of China Department of Biotechnology and Biomedicine, DTUBioengineering, Technical University of Denmark, 2800 Kongens Lyngby, Denmark 17 8 Department of Health Sciences, Faculty of Science, University of Mauritius, Reduit, Mauritius Botanischer Garten und Botanisches Museum Berlin, Freie Universität Berlin, Königin-Luise-Straße 6–8, 14195 Berlin, Germany 9 Formerly Department of Botany, Goa University, Goa, India 123 Fungal Diversity (2019) 96:1–242 18 Laboratório de Botânica/Liquenologia, Instituto de Biociências, Universidade Federal de Mato Grosso do Sul, Avenida Costa e Silva s/n, Bairro Universitário, Campo Grande, Matto Grosso do Sul CEP 79070-900, Brazil 19 Departamento de Biociências, Universidade Federal de Sergipe, Itabaiana, Sergipe CEP: 49500-000, Brazil 20 21 22 23 24 25 26 27 28 36 Yunnan Institute of Food Safety, Kunming University of Science and Technology, Kunming 650500, People’s Republic of China 37 Guizhou Key Laboratory of Agricultural Biotechnology, Guizhou Academy of Agricultural Sciences, Guiyang 550006, Guizhou, People’s Republic of China Shenzhen Key Laboratory of Microbial Genetic Engineering, College of Life Sciences and Oceanography, Shenzhen University, Shenzhen 518060, People’s Republic of China 38 Engineering Research Center of Southwest BioPharmaceutical Resources, Ministry of Education, Guizhou University, Guiyang 550025, People’s Republic of China Industrial Science and Management (International Program), Faculty of Science and Technology, Thammasat University (Rangsit Center), Klong Luang, Rangsit, Pathumthani 12121, Thailand 39 Jodrell Laboratory, Royal Botanic Gardens, Kew, Surrey TW9 3AB, UK 40 Institute of Microbiology, Beijing Forestry University, PO Box 61, Beijing 100083, People’s Republic of China 41 Shandong Provincial Key Laboratory of Applied Mycology, Qingdao Agricultural University, Qingdao 266109, People’s Republic of China 42 Rud̄er Bošković Institute, Bijenička cesta 54, 10000 Zagreb, Croatia 43 National Fungal Culture Collection of India (NFCCI), Biodiversity and Palaeobiology Group, MACS Agharkar Research Institute, G.G. Agarkar Road, Pune 411 004, India 44 State Key Laboratory of Mycology, Institute of Microbiology, Chinese Academy of Sciences, Beijing 100101, People’s Republic of China 45 Institute of Botany, Czech Academy of Sciences, 252 43 Průhonice, Czech Republic 46 Las Muros, 09420 Rimont, France 47 Institute for Clinical Biochemistry and Diagnostics, University Hospital Hradec Králové, 500 05 Hradec Králové, Czech Republic 48 Centro de Biociências, Departamento de Micologia, Universidade Federal de Pernambuco, Avenida da Engenharia, S/N 50740-600 – Cidade Universitária, Recife, Pernambuco, Brazil 49 Department Research, Meise Botanic Garden, Nieuwelaan 38, 1860 Meise, Belgium 50 Fédération Wallonie-Bruxelles, Direction Générale de l’Enseignement non obligatoire et de la Recherche scientifique, Rue A. Lavallée 1, 1080 Brussels, Belgium 51 Faculdade de Ciências, Biosystems and Integrative Sciences Institute (BioISI), Universidade de Lisboa, Campo Grande, 1749-016 Lisbon, Portugal 52 A.M.B, Circolo Micologico ‘‘Giovanni Carini’’, C.P. 314, Brescia, Italy 53 A.M.B. Gruppo, Micologico Forlivese ‘‘Antonio Cicognani’’, Via Roma 18, Forlı́, Italy 54 Società per gli Studi Naturalistici della Romagna, C.P. 143, Bagnacavallo, RA, Italy 55 Russian Research Institute of Floriculture and Subtropical Crops, Yana Fabritsiusa Street 2/28, Sochi, Krasnodar region, Russia 354002 56 Department of Botany and Microbiology, College of Science, King Saud University, P.O Box 2455, Riyadh 11451, Kingdom of Saudi Arabia Instituto de Ecologı́a, A. C., Apartado Postal 63, 91000 Xalapa, Veracruz, Mexico College of Agriculture and Biological Science, Dali University, Dali 671003, Yunnan, People’s Republic of China Departamento Fitopatologia, Universidade Federal de Viçosa, Viçosa, Minas Gerais, Brazil Universidade Federal de Viçosa, Departamento de Microbiologia, Viçosa, Minas Gerais, Brazil Institut de Systématique, Evolution, Biodiversité (ISYEB – UMR 7205), Muséum national d’Histoire naturelle, Sorbonne Université, CNRS, CP 39, 12 Rue Buffon, 75005 Paris, France Centre National de Recherche sur l’Environnement (CNRE), Lab. de Microbiologie de l’Environnement (LME), BP 1739, Antananarivo, Madagascar Department of Plant Protection, Agroscope ChanginsWädenswil Research Station ACW, Rte De Duiller, 1260 Nyon 1, Switzerland 29 Departamento de Micologia, Universidade Federal de Pernambuco, Avenida da Engenharia, S/N, Recife 50740-600, Brazil 30 Museu Paraense Emı́lio Goeldi, Coord. Botânica, Av. Perimetral 1901, Terra Firme, C.P. 399, Belém, PA 66040-170, Brazil 31 32 33 34 35 241 Instituto de Ciências Biológicas, Laboratório de Microbiologia, Av. Augusto Corrêa, Belém, Pará 66075-110, Brazil Coordenação de Engenharia Ambiental, Universidade do Estado do Amapá, Av. Presidente Vargas, Macapá 68900-070, Brazil Programa de Pós-Graduação em Biologia de Fungos, Departamento de Micologia Prof. Chaves Batista, Centro de Biociências, Universidade Federal de Pernambuco, Recife, PE 50670-901, Brazil Department of Entomology and Plant Pathology, Faculty of Agriculture, Chiang Mai University, Chiang Mai 50200, Thailand Beijing Key Laboratory of Environment Friendly Management on Fruit Diseases and Pests in North China, Institute of Plant and Environment Protection, Beijing Academy of Agriculture and Forestry Sciences, Beijing 100097, People’s Republic of China 123 242 Fungal Diversity (2019) 96:1–242 57 Biodiversity Integration Knowledge Center, Arizona State University, PO Box 874108, Tempe, AZ 85287-4108, USA 65 Royal Botanic Gardens, Kew, Richmond, Surrey TW9 3DS, UK 58 Charles Darwin Foundation for the Galapagos Islands, Puerto Ayora, Ecuador 66 59 Instituto Nacional de Biodiversidad, Quito, Ecuador School of Life Science and Technology, University of Electronic Science and Technology of China, Chengdu 611731, People’s Republic of China 60 67 College of Science and Engineering, Flinders University, GPO Box 2100, Adelaide, SA 5001, Australia Institute of Animal Science, State Key Laboratory of Animal Nutrition, Chinese Academy of Agricultural Sciences, Beijing 100193, People’s Republic of China 61 College of Life Sciences, Hebei University, Baoding 071002, Hebei Province, People’s Republic of China 68 Center of Excellence in Microbial Diversity and Sustainable Utilization, Chiang Mai University, Chiang Mai 50200, Thailand 69 Academy of Science, The Royal Society of Thailand, Bangkok 10300, Thailand 70 Institute of Applied Ecology, Chinese Academy of Sciences, Shenyang 110016, People’s Republic of China 62 State Key Laboratory Breeding Base of Green Pesticide and Agricultural Bioengineering, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Guizhou University, Guiyang 550025, People’s Republic of China 63 Laboratory of Mycology, Institute of Botany, Academy of Sciences of the Republic of Uzbekistan, 32 Durmon Yuli Street, Tashkent, Uzbekistan 100125 71 Centre for Mountain Futures (CMF), Kunming Institute of Botany, Kunming 650201, Yunnan, People’s Republic of China 64 Department of Ecology, University of Kassel, HeinrichPlettStrasse, 40, 34132 Kassel, Germany 72 Center of Excellence in Bioresources for Agriculture, Industry and Medicine, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand 123